effects of light, co2 and reactor design on growth of algae

185
Effects of Light, Co 2 and Reactor Design on Growth of Algae: An Experimental Approach to Increase Biomass Production by Sebastián Mejía Rendón A Thesis presented to The University of Guelph In partial fulfilment of requirements for the degree of Doctor of Philosophy In Land Resources Sciences Guelph, Ontario, Canada © Sebastián Mejía Rendón, April, 2014

Upload: others

Post on 11-Sep-2021

4 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Effects of Light, Co2 and Reactor Design on Growth of Algae

Effects of Light, Co2 and Reactor Design on Growth of Algae: An

Experimental Approach to Increase Biomass Production

by

Sebastián Mejía Rendón

A Thesis

presented to

The University of Guelph

In partial fulfilment of requirements

for the degree of

Doctor of Philosophy

In

Land Resources Sciences

Guelph, Ontario, Canada

© Sebastián Mejía Rendón, April, 2014

Page 2: Effects of Light, Co2 and Reactor Design on Growth of Algae

ABSTRACT

Effects of Light, Co2 and Reactor Design on Growth of Algae: An Experimental Approach

to Increase Biomass Production

 

Sebastian Mejia Rendon Advisor:

University of Guelph, 2014 Paul Voroney

 

The increase in atmospheric carbon dioxide concentrations and resulting global climate change

coupled with the high cost of fossil fuels are encouraging the search for alternative sources of

fuels and methods to produce food without using agricultural land. Algal biomass production has

the potential to address these problems; algae are rich in fats, proteins and carbohydrates. Algae

fix carbon dioxide and are relatively easy to grow. They can be grown in both fresh and marine

water. Algal oil can be used as a feedstock to make biodiesel, the omegas and antioxidants can

be used as food supplements and components of medicines; the proteins can be used for animal

feedstock and food suppliments; and the carbohydrates can be used to produce ethanol. Study of

the factors affecting growth of microalgae play a critical role for the development of efficient

methodologies for harvesting their products. Experiments were conducted with six algae to

evaluate effects of light source, CO2 concentration and reactor design; small bubble column

versus flat panel photo-bioreator.

Page 3: Effects of Light, Co2 and Reactor Design on Growth of Algae

Light treatments that supplemented artificial light with daily solar radiation at 4614 mg L-1

resulted in 15% more biomass production than those treatments that provided 24 h of artificial

illumination (4016 mg L-1) during 49 days of incubation.Treatments with Chlorella vulgaris

during 15 d that combined continuous exposure from LEDs emitting blue + red radiation (430

and 625 nm) and from white light resulted in greater biomass production, 1003 mg L-1 and 1045

mg L-1, respectively, than those from a single light source (red at 571 mg L-1 and blue at 828 mg

L-1). Growth of Nannochloris sp and Scenedesmus sp during 13 d of incuabion exposed to

flashing lights in a bubble column photo-bioreactor resulted in more biomass production (1041

mg L-1 and 1130 mg L-1, respectively) and higher cell weight than those algae growing under

continuous light (892 mg L-1 and 1018 mg L-1, respectively. Experiments with bubble column

photo-bioreactors and Chlorella vulgaris growing under increasing CO2 concentrations, from

0.035% to 8.5%, during a 15 d incubation increased biomass production from 148 mg L-1 to

1592 mg L-1. A flat panel, industrial scale photo-bioreactor resulted in biomass production of

2337 mg L-1 during 15 d, which was significantly more than that obtained in bubble column

reactors.

This research has shown that algal biomass production can be increased when grown under

illumination that supplemented solar radiation with artificial light during the night time. An

increase in the concentration of CO2 supplied to the algae increased algal biomass production 10

fold. Exposure as light dark cycles, the flashing effect, which utilitzes 85-90% less energy

consumption than exposure to continuous light, resulted in greater biomass productiont. A light

path less than 5 cm resulted in increased biomass production by >1.5 fold. Algae growing in a

flat panel bioreactor developed the highest biomass production, and the research suggests that

this bioreactor design can be a reliable system for growing algae at a large scale.

Page 4: Effects of Light, Co2 and Reactor Design on Growth of Algae

iv 

 

ACKNOWLEDGEMENTS

I would like to thank all of those who have been involved during the course of this research

study, especially to Gabriel Jaime Colmenares who assisted in setting up and conducting the

algal growth experiments. I have special thanks to my parents who have provided support and

encouragement during my studies. I wish to thank the staff in the former Department of Land

resource Science (now the School of the Environmental Sciences) for their assistance. I also

would like to thank the members of my PhD advisory committee: Drs. Kari Dunfield, Hongde

Zhou and Paul Voroney, for their guidance through out my program. I have a very special thanks

to my supervisor Paul Voroney who provided me the opportunity to conduct research in the area

of photosynthetic microorganisms and guided my PhD program.

Page 5: Effects of Light, Co2 and Reactor Design on Growth of Algae

 

 

TABLE OF CONTENTS

CHAPTER 1 INTRODUCTION _______________________________________________ 1

1.1 BACKGROUND ____________________________________________________________ 1 1.1.1 ALGAL GROWTH FOR PHARMACEUTICAL PRODUCTION AND FOOD SUPPLEMENTS ___ 1 1.1.2 OIL SYNTHESIS ________________________________________________________________ 2 1.1.3 CO2 FIXATION ___________________________________________________________________ 2 1.1.4 PHOTOBIOREACTORS __________________________________________________________ 3 1.1.5 ILLUMINATION ________________________________________________________________ 4

1.1.5.1 FLASHING LIGHT AND DARK PERIOD ____________________________________________ 4 1.1.5.2 PHOTOINHIBITION IN ALGAE _________________________________________________ 7 1.1.5.3 WEIGHT LOSS DURING NIGHT PERIOD ________________________________________ 9

1.1.6 ALGAE GROWTH UNDER ARTIFICIAL AND OUTDOOR CONDITIONS ________________ 9 1.1.6.1 GROWTH UNDER INDOOR ARTIFICIAL CONDITIONS ____________________________ 9 1.1.6.2 GROWTH UNDER OUTDOOR CONDITIONS ____________________________________ 10

1.1.7 COMMERCIAL PRODUCTION OF ALGAE ________________________________________ 10 1.1.8 ALGAE USED DURING THE COURSE OF THIS RESEARCH _________________________ 11

1.2 FORMAT OF THE THESIS ____________________________________________________ 12

1.3 REFERENCES ____________________________________________________________ 14

CHAPTER 2 RESEARCH OBJECTIVES ______________________________________ 24

2.1 GENERAL GOAL _________________________________________________________ 24

2.2 SPECIFIC RESEARCH OBJECTIVES _______________________________________ 24

CHAPTER 3 THE PHOTOSYNTHETIC MACHINERY: A DETAILED DESCRIPTION OF THE LIGHT REACTION; PSI, PSII, ATP SYNTHASE, STATE OF TRANSITION, PHOTO-DAMAGE, REGENERATION, AND MACHINERY ARCHITECTURE ________ 25

3.1 ABSTRACT _______________________________________________________________ 25

3.2 INTRODUCTION __________________________________________________________ 26

3.3 EVOLUTION OF PHOTOSYNTHESIS: ANOXYGENIC AND OXYGENIC ________ 28 3.3.1 ANOXYGENIC PHOTOSYNTHESIS ________________________________________________ 30

3.3.1.1 THE PURPLE NON-SULFUR BACTERIA _______________________________________ 30 3.3.1.2 THE GREEN AND PURPLE SULFUR BACTERIA _________________________________ 31

Page 6: Effects of Light, Co2 and Reactor Design on Growth of Algae

vi 

 

3.3.1.3 HELIOBACTERIA ___________________________________________________________ 31 3.3.1.4 RHODOPSIN PHOTOSYNTHESIS ______________________________________________ 31

3.3.2 OXYGENIC PHOTOSYNTHESIS _________________________________________________ 32 3.3.3 ARCHITECTURE OF THE PHOTOSYNTHETIC APPARATUS IN PLANTS AND ALGAE __ 33 3.3.4 GENERAL DESCRIPTION AND FUNCTION OF PHOTOSYSTEM II (PSII) ______________ 36

3.3.4.1 PSII REACTION CENTER PROCESSES: P680, D1-D2 PROTEINS, PHEOPHYTIN, OXYGEN EVOLUTION CENTER (OEC), WATER OXIDATION PROCESS (WOC) AND PLASTOQUINONES __________________________________________________________ 39 

3.3.4.2 OXYGEN EVOLUTION CENTER (OEC) ____________________________________________ 41 

3.3.4.3  ARCHITECTURE AND DYNAMICS OF THE WATER OXIDATION PROCESS (WOC) __ 42 

3.3.4.4  MECHANISM OF ENERGY TRANSFER FROM PSII TO THE MOBILE CARRIERS  ____ 43 

3.3.4.5   ROLE OF PLASTOQUINONES IN PSII (MOBILE CARRIERS)  ______________________ 44 

3.3.4.6   DAMAGE AND REGENERATION OF PSII _______________________________________ 45 

3.3.4.7   ROLE OF PSB27 IN THE REGENERATION OF PSII _______________________________ 46 

3.3.5   GENERAL ARCHITECTURE AND FUNCTION OF PSI _______________________________ 47 

3.3.5.1   PSI REACTION CENTER CHLOROPHYLL P700 __________________________________ 48 

3.3.5.2   REDOX POTENTIAL OF PSI ___________________________________________________ 50 

3.3.6   MECHANISMS TO MINIMIZE PHOTO-DAMAGE IN PSI AND PSII ___________________ 51 3.3.7   STATE TRANSITION PSII AND PSI SYNTHESIS RATIO AND PHOTOSYNTHETIC

ACTIVITY  ____________________________________________________________________ 51 

3.3.8   BACTERIA STATE OF TRANSITION _____________________________________________ 53 

3.3.9   CONCLUSIONS ________________________________________________________________ 54 

3.4   REFERENCES _____________________________________________________________  55 

 

CHAPTER 4   GROWTH OF TWO MARINE ALGAE AND TWO FRESH WATER ALGAE UNDER SOLAR AND LED ILLUMINATION _____________________________  60 

4.1   ABSTRACT  _______________________________________________________________  60 

4.2   INTRODUCTION __________________________________________________________  60 

4.3   MATERIALS AND METHODS  ______________________________________________  62 4.3.1   MICROORGANISM  ____________________________________________________________ 62 

4.3.2   LIGHT SOURCE  _______________________________________________________________ 63 

4.3.2.1   EXPERIMENT 1 (DAILY SOLAR RADIATION) ___________________________________ 63 

4.3.2.2   EXPERIMENT 2 (DAILY SOLAR PLUS 12 H WHITE LIGHT) _______________________ 63 

4.3.2.3   EXPERIMENT 3 (24 H BLUE LEDS)  ____________________________________________ 63 

4.3.2.4   EXPERIMENT 4 (24 H RED + BLUE LEDS) ______________________________________ 63 

4.3.2.5   EXPERIMENT 5 (DAILY SOLAR PLUS 12H RED AND BLUE LEDS) ________________ 64 

4.3.3  LIGHT DISPERSION ______________________________________________________________ 64 

4.3.4   MIXING, QUANTIFYING CARBON SOURCE UNDER AUTOTROPHIC GROWTH AND TEMPERATURE MEASUREMENTS ______________________________________________ 65 

4.3.5 GROWTH MEDIA PREPARATION ____________________________________________________ 65 

4.3.6   CELL DENSITY ANALYSIS _____________________________________________________ 66 

4.3.7   GROWTH RATE METHODOLOGY _______________________________________________ 66 

Page 7: Effects of Light, Co2 and Reactor Design on Growth of Algae

vii 

 

4.3.8  BIOMASS AND CELL WEIGHT METHODOLOGY ____________________________________ 66 

4.3.9  EXPERIMENTAL DESIGN AND STATISTICAL ANALYSIS ____________________________ 67 

4.4   RESULTS AND DISCUSSION  _______________________________________________  67 

4.4.1   ALGAL GROWTH UNDER DAILY SOLAR RADIATION _____________________________ 68 

4.4.2   ALGAL GROWTH UNDER DAILY SOLAR RADIATION + 12 H WHITE LIGHT _________ 68 

4.4.4   ALGAL GROWTH UNDER 24 H BLUE + RED LEDS ________________________________ 69 

4.4.5   ALGAL GROWTH UNDER DAILY SOLAR RADIATION + 12 H BLUE +RED LEDS ______ 70 4.4.6   FINAL POPULATION DENSITY FOR FOUR ALGAL SPECIES GROWING UNDER 5 LIGHT

TREATMENT __________________________________________________________________ 70 

4.4.7   SPECIFIC GROWTH RATE FOR VARIOUS LIGHT TREATMENTS ____________________ 72 

4.4.8   BIOMASS PRODUCTION  _______________________________________________________ 75 

4.4.9  CELL WEIGHT ___________________________________________________________________ 76 

4.5   CONCLUSIONS  ___________________________________________________________  77 

4.6   REFERENCES _____________________________________________________________  79 

 

CHAPTER 5   BIOMASS PRODUCTION OF SIX ALGA SPECIES GROWING UNDER LED AND SOLAR RADIATION _______________________________________________  85 

5.1   ABSTRACT  _______________________________________________________________  85 

5.2   INTRODUCTION __________________________________________________________  85 

5.3   MATERIALS AND METHODS  ______________________________________________  88 5.3.1   MICROORGANISM  ____________________________________________________________ 88 

5.3.2   BIOREACTOR SYSTEM _________________________________________________________ 88 

5.3.3   LIGHT SOURCE  _______________________________________________________________ 88 

5.3.4  LIGHT DISPERSION ______________________________________________________________ 89 

5.3.5   EXPERIMENT 1: 24 H RED + BLUE LEDS _________________________________________ 89 

5.3.6   EXPERIMENT 2: DAILY SOLAR + 12 H LEDS ______________________________________ 89 5.3.7   MIXING, QUANTIFYING CARBON SOURCE UNDER AUTOTROPHIC GROWTH AND

TEMPERATURE MEASUREMENTS  ______________________________________________ 90 

5.3.8   CULTURE MEDIUM ____________________________________________________________ 90 

5.3.9   BIOMASS DETERMINATION ____________________________________________________ 90 

5.3.10   EXPERIMENTAL DESIGN AND STATISTICAL ANALYSIS __________________________ 91 

5.4   RESULTS _________________________________________________________________  91 5.4.1   BIOMASS PRODUCTION FOR FRESH WATER ALGAE GROWING UNDER TWO

DIFFERENT LIGHT TREATMENTS _______________________________________________ 91 

5.4.2   MEAN AND TOTAL BIOMASS PRODUCTION _____________________________________ 94 

5.4.3   BIOMASS VARIANCE __________________________________________________________ 96 

5.5.   DISCUSSION AND CONCLUSIONS __________________________________________  97 

5.6  REFERENCES ____________________________________________________________ 100 

Page 8: Effects of Light, Co2 and Reactor Design on Growth of Algae

viii 

 

 

CHAPTER 6   EFFECT OF USING DIFFERENT LEDS LIGHTS ON THE GROWTH AND BIOMASS PRODUCTION OF CHLORELLA VULGARIS ____________________  105 

6.1   ABSTRACT  ______________________________________________________________ 105 

6.2   INTRODUCTION _________________________________________________________ 105 

6.3   MATERIALS AND METHODS  _____________________________________________ 107 6.3.1   MICROORGANISM  ___________________________________________________________ 107 

6.3.2   LIGHT SOURCE  ______________________________________________________________ 107 

6.3.3   LIGHT EMISSION INTENSITY __________________________________________________ 107 

6.3.4   REACTOR MANAGEMENT  ____________________________________________________ 109 

6.3.5   CULTURE MEDIUM ___________________________________________________________ 109 

6.3.6   EXPERIMENTAL DESIGN AND STATISTICAL ANALYSIS _________________________ 109 

6.3.7   ALGAL POPULATION, LIGHT ABSORPTION PEAKS AND BIOMASS ANALYSIS _____ 110 

6.4.   RESULTS AND DISCUSSION  ______________________________________________ 112 

6.4.1   THE EFFECT LIGHT IN THE ALGAL ABSORPTION SPECTRA ______________________ 112 

6.4.2. POPULATION GROWTH OF C. VULGARIS EXPOSED TO DIFFERENT LIGHT SOURCES  ___ 113 

6.4.3.3 STANDING BIOMASS PRODUCTION UNDER DIFFERENT LIGHT SOURCES  _________ 115 

6.4.3.3.1 STANDING BIOMASS  _____________________________________________________ 115 

6.4.3.4 EFFECT OF LIGHT SOURCE ON ALGAL CELL WEIGHT  ___________________________ 117 

6.5   CONCLUSIONS  __________________________________________________________ 118 

6.6   REFERENCES ____________________________________________________________ 119 

 

CHAPTER 7   BIOMASS PRODUCTION OF MARINE AND FRESH WATER ALGA GROWING UNDER FLASHING AND CONTINUOUS LED LIGHT ________________  125 

7.1   ABSTRACT  ______________________________________________________________ 125 

7.2   INTRODUCTION _________________________________________________________ 125 

7.3   MATERIALS AND METHODS  _____________________________________________ 127 7.3.1   MICROORGANISM  ___________________________________________________________ 127 

7.3.2   PHOTO REACTOR ____________________________________________________________ 127 

7.3.3   LIGHT SOURCE AND FLASHING DEVICE _______________________________________ 128 

7.3.4   LIGHT DISPERSION ___________________________________________________________ 128 

7.3.5   MIXING, QUANTIFYING CARBON SOURCE UNDER AUTOTROPHIC GROWTH AND TEMPERATURE MEASUREMENTS  _____________________________________________ 129 

7.4.5   CULTURE MEDIUM ___________________________________________________________ 129 

7.4.6   BIOMASS DETERMINATION ___________________________________________________ 129 

7.4.7   EXPERIMENTAL DESIGN AND STATISTICAL ANALYSIS _________________________ 130 

7.5   RESULTS AND DISCUSSION  ______________________________________________ 131 

Page 9: Effects of Light, Co2 and Reactor Design on Growth of Algae

ix 

 

7.5.1   STANDING BIOMASS UNDER DIFFERENT LIGHT ________________________________ 131 

7.5.2   DAILY BIOMASS PRODUCTION ________________________________________________ 132 

7.5.3   CHANGES IN POPULATION DENSITY  __________________________________________ 132 

7.5.4   CELL WEIGHT  _______________________________________________________________ 134 

7.6   DISCUSSION AND CONCLUSIONS _________________________________________ 136 

7.8   REFERENCES ____________________________________________________________ 138 

 

CHAPTER 8   EFFECT OF CARBON DIOXIDE CONCENTRATION ON THE GROWTH RESPONSE OF CHLORELLA VULGARIS UNDER FOUR DIFFERENT LED ILUMINATION  142 

8.1   ABSTRACT  ______________________________________________________________ 142 

8.2   INTRODUCTION _________________________________________________________ 142 

8.3   MATERIALS AND METHODS  _____________________________________________ 143 8.3.1   MICROORGANISM  ___________________________________________________________ 143 

8.3.2   LIGHT SOURCE  ______________________________________________________________ 143 

8.3.3   LIGHT DISPERSION ___________________________________________________________ 144 

8.3.4   PHOTOBIOREACTOR AND CO2 SUPPLY  ________________________________________ 145 

8.3.5   CULTURE MEDIUM __________________________________________________________ 145 

8.3.6   EXPERIMENTAL DESIGN AND STATISTICAL ANALYSIS _________________________ 146 

8.3.7   BIOMASS ANALYSIS  _________________________________________________________ 146 

8.3.8. STATISTICAL ANALYSIS ______________________________________________________ 146 

8.4   RESULTS AND DISCUSSION  ______________________________________________ 147 

8.4.1   STANDING BIOMASS PRODUCTION UNDER INCREASING CO2 CONCENTRATIONS _ 147 

8.4.2   STATISTICAL ANALYSIS OF THE STANDING BIOMASS __________________________ 148 

8.4.3 LIGHT SOURCE EFFECTS ON BIOMASS PRODUCTION ___________________________ 150 

8.5. CONCLUSIONS _____________________________________________________________ 151 

8.6   REFERENCES ____________________________________________________________ 153 

 

CHAPTER 9   BIOMASS PRODUCTION OF TWO MARINE AND TWO FRESH WATER ALGAE GROWING IN AN INDUSTRIAL-SCALE THIN FLAT PANEL PHOTO-BIOREACTOR UNDER FOUR CARBON DIOXIDE CONCENTRATIONS ___________  157 

9.1   ABSTRACT  ______________________________________________________________ 157 

9.2   INTRODUCTION _________________________________________________________ 157 

9.3   MATERIALS AND METHODS  _____________________________________________ 159 9.3.1   MICROORGANISM  ___________________________________________________________ 159 

9.3.2   FLAT PANEL REACTOR _______________________________________________________ 159 

Page 10: Effects of Light, Co2 and Reactor Design on Growth of Algae

 

9.3.3   ILLUMINATION AND TEMPERATURE __________________________________________ 160 

9.3.4   CARBON DIOXIDE SUPPLY ____________________________________________________ 161 

9.3.5   CULTURE MEDIUM ___________________________________________________________ 161 

9.3.6   BIOMASS DETERMINATION ___________________________________________________ 162 

9.3.7   EXPERIMENTAL DESIGN AND STATISTICAL ANALYSIS _________________________ 162 

9.4   RESULTS ________________________________________________________________ 163 9.4.1   CHANGES IN ALGAL CELL DENSITY UNDER AIR SUPPLY (380-410 PPM OF CO2)  ___ 163 

9.4.2   BIOMASS PRODUCTION WITH INCREASING CO2 TREATMENTS  __________________ 164 9.4.3   BIOMASS COMPARISON OF ALGAE GROWING AT DIFFERENT CO2 CONCENTRATIONS     _____________________________________________________________________________ 165 

9.5   DISCUSSION AND CONCLUSION __________________________________________ 166 

9.6   REFERENCES ____________________________________________________________ 167 

CHAPTER 10:   GENERAL CONCLUSIONS __________________________________  169  

Page 11: Effects of Light, Co2 and Reactor Design on Growth of Algae

xi 

 

 

LIST OF TABLES

Table 1.1. Algae biomass production in different growing systems  ______________________  4 

 

Table 1.2. Studies with algae under flashing lights  ___________________________________  6 

 

Table 1.3. Optimal light intensity for growing algae __________________________________  7 

 

Table 1.4. Quantum yield in photon absorption by biomass under different photon flux _____  8 

 

Table 1.5. Taxonomy of the six algal species grown during the course of this research ______  11 

 

Table 4.1. Population densities (cells mL‐1) after 18 days of incubation for four different algae 

under five different light treatments _____________________________________________  71 

 

Table 4.2. Growth rate (d‐1) under different lights treatments after 18 d of incubation  _____  73 

 

Table 5.1. Biomass production for 6 different alga species ____________________________  87 

 

Table 5.2. Biomass production for three fresh water algae under two light sorces _________  93 

 

Table 5.3. Biomass production for three marine algae under two light sources ____________  93 

 

Table 5.4. ANOVA test for total biomass production _________________________________  96 

 

Table 6.1. Effect of distance from the LED light source on light power __________________  108 

 

Table 6.2. Standing biomass of C. vulgaris growing under white, blue, red and red + blue light

__________________________________________________________________________  115 

Page 12: Effects of Light, Co2 and Reactor Design on Growth of Algae

xii 

 

Table 6.4. Cell weight of C. vulgaris grown under 4 different light sources  ______________  117 

 

Table 7.1. Cell weight of two algae species under two light treatments _________________  135 

 

Table 8.1. LED light dispersion and light reducing intensity through distance ____________  145 

 

Table 8.2. Analysis of variance of the standing biomass for four different CO2 concentration 

under costant light sources  ___________________________________________________  149 

 

Table 8.3. Analyses of the standing biomass if the light regime changes and the CO2 

concentration remains constant ________________________________________________  149 

 

Table 8.4. Mean standing biomass production under increasing CO2 concentrations and 

exposed to four different light sources  __________________________________________  151 

Page 13: Effects of Light, Co2 and Reactor Design on Growth of Algae

xiii 

 

 

LIST OF FIGURES

Figure 1.1. Features of the six algal species studied during the course of this research        12 

Figure 3.1. Chloroplast structure ________________________________________________  34 

Figure 3.2. Thylakoid disk including PSI, PSII and ATP synthase  ________________________  35 

Figure 3.3. Organization of the thylakoid disk including the stroma lamellae  _____________  36 

Figure 3.4. Three‐dimensional structure of the PSII antennae and reaction center _________  38 

Figure 3.5. Proposed structure of PSII based on X‐ray images  _________________________  40 

Figure 3.6. Photosystem I structure ______________________________________________  49 

Figure 3.7. Photosynthetic apparatus in algae ______________________________________  53 

Figure 4.1. Algal growth under daily solar radiation _________________________________  68 

Figure 4.2. Algal growth under daily solar radiation + 12 h white light ___________________  68 

Figure 4.3. Algal growth under to 24 h blue LEDs in 4‐L reactors _______________________  69 

Figure 4.4. Algal growth under 24 h red + blue LEDs _________________________________  70 

Figure 4.5. Algal growth under daily solar radiation + 12 h blue +red LEDs  _______________  70 

Figure 5.1. Mean algal biomass production under two light treatments  _________________  94 

Figure 5.2. Total algal biomass production under two light treatments __________________  95 

Figure 5.3. Biomass variance during the experiment period ___________________________  97 

Page 14: Effects of Light, Co2 and Reactor Design on Growth of Algae

xiv 

 

Figure. 6.1. Experiment set up, including 24 replicates ______________________________  110 

Figure. 6.2. Algal absorption spectra after 15‐d incubation exposed to blue, red, red+blue and white LED lights _____________________________________________________________  112  

Figure. 6.3. Growth curves of the four light treatments with 6 replicates of each treatment  114 

Figure 7.1. Standing biomass production under two light treatments for a marine algal Nannochloris sp and a fresh water alga Scenedesmus sp ____________________________  131  

Figure 7.2. Population density of two algae under two light treatments ________________  133 

Figure 7.3. Photograph analysis of Nannochloris sp and Chlorella sp include size and some organelles _________________________________________________________________  134  

Figure 8.1. Changes in standing biomass under different concentration of CO2 and different wavelength illumination ______________________________________________________  148  

Figure 9.1. Reactor unit from experimental set up _________________________________  160 

Figure 9.2. Experimental set up showing the night illumination _______________________  161 

Figure 9.3. Population density of four alga species growing in a flat panel industrial scale reactor__________________________________________________________________________  163  

Figure 9.4. Biomass production of two fresh water algae C. vulgaris, Scenedesmus sp and two marine algae Tetraselmis sp, Nannochloris sp exposed to increasing CO2 concentrations in an FPB  ______________________________________________________________________  164   

Page 15: Effects of Light, Co2 and Reactor Design on Growth of Algae

 

 

CHAPTER 1 INTRODUCTION

1.1 BACKGROUND

GLOSARY

Light intensivity: 1 µmol m-2 s-1 = 1 µE m-2 s-1 = 0.342 W m-2 = 54 Lux

Standing biomass: The weight of the algae population in a volume unit (mg L-1 or g L-1)

Biomass production: It is the weight of the algae that is been produce in a period of time in a volume unit, sometimes the standing biomass is the biomass production over time (mg L-1 d-1)

Flashing light: Periods of illumination and periods of darkness in milliseconds

1.1.1 ALGAL GROWTH FOR PHARMACEUTICAL PRODUCTION AND FOOD SUPPLEMENTS

Algal biomass production has great potential as a source of pharmaceutical metabolites and

biochemical compounds such as antibiotics, neutraceuticals and other natural drugs, for treating

diseases in humans and animals (Fenical, 1997; Wright et al., 1999), and for use in different food

supplements (Fan et al., 2008; Lucumi and Posten, 2006; Hu et al., 1996a). For example, the

marine algae Nannochloropsis sp is known for the production of astaxanthin, zeaxanthin and

canthaxanthin as well as for production of useful lipids and biomass with high nutritional value

for human consumption (Rocha et al., 2003). The company Quantum Nutrition Inc located in

Ontario Canada produces a food supplement called Schinoussa Sea Vegetables, which is a

mixture of algal species suchs as Spirulina sp, Chlorella sp AFA and red algae species. The use

of photosynthetic microorganisms for the production of pharmaceutical compounds has been

called molecular farming (Lucumi and Posten, 2006).

The algal Haematococcus pluvialis has been grown for the production of astaxanthin (3,3-

dihydroxy-β, β-carotene-4, 4-dione) a pharmaceutical compound and a food supplement found in

Page 16: Effects of Light, Co2 and Reactor Design on Growth of Algae

 

stores under different brands such as Now Foods a company located in Bloomingdale Illinois

USA. For the production of asthaxanthin H. pluviais is grown under lumostatic conditions (light

intensity 40 µmol m-2 s-1 and air containing 5% CO2) a standing biomass of 5 g L-1, a cell

population 6*106 cells mL-1 and a astaxanthin production of 130 mg L-1 was obtained (Lee et al.,

2006). Using a flat panel photobioreactor with superficial mixing velocity of 0.4 cm s-1 and light

intensity of 20 µmol m-2 s-1, Issarapayup et al. (2009) obtained a population of 4.1 *105 cells mL-

1 and a population doubling time of 46 h.

1.1.2 OIL SYNTHESIS

The growth of algal species for lipid production is an alternative technology for use as a

renewable energy source. However, to achieve this goal it is essential that current algal lipid

contents be increased. A technique for increasing the lipid content of alga is growth of under low

nitrogen supply. Under nitrogen limitation, Chlorella vulgaris increased lipid synthesis from

17% to 58%, but the biomass production was reduced from 680 mg L-1 d -1 to 24 mg L-1 d-1

(Doucha and Livansky, 2009; Scragg et al., 2002; Watanabe and Saiki, 1997). While algae store

carbohydrates synthesized during the period of photosnthesis (exposure to light) and consume

those carbohydrates during the dark period with respiration, fatty acid biosynthesis and

consumption occur during the light period (Rebolloso Fuentes et al., 1999). The composition of

algal fatty acids has been shown to be affected by the conditions under which they are grown,

and for the algal species Pavlova lutheri the fatty acids ranged in size from 14 to 22 carbon

chains (Meireles et al., 2008).

1.1.3 CO2 FIXATION

Algae cells contain 45-50% organic carbon (on dry weight basis), therefore it is the highest

metabolic requirement constraining growth after water. The carbon is required in the form of

Page 17: Effects of Light, Co2 and Reactor Design on Growth of Algae

 

CO2 for photosynthesis reactions, and theoretically the production of 1 kg of algal biomass

requires the uptake of 1.83 kg CO2 (Doucha et al., 2005). In reality, algal cells also produce

various carbon compounds as sub-products that are released from the cell during growth. Thus,

the amount of CO2 required for the production of 1 kg of algae biomass is 4.4 kg, taking into

consideration a net CO2 absorption of 38.7% growing under reactor conditions (Doucha et al.,

2005).

1.1.4 PHOTOBIOREACTORS

Horizontal tubular photobioreactors are the most practical and scalable systems for algal

production and are commonly used for pharmaceutical and industrial purposes. For a tubular

photobioreactor it is recommended that it be 2.4-10 cm in diameter and have a maximum length

of 80-84 m, and have a nutrient media flow velocity of 30-50 cm s-1 (Acién Fernández et al.,

2001; Molina Grima et al., 1999; Baquerisse et al., 1999). Bubble-column systems are also a

common design for photobioreactors, with recommended dimensions being 20 cm in diameter

and 4 m column height (Sanchez Miron et al., 1999). Both systems can provide sustainable

conditions for growing algal species. In a bubble column photobioreactor the marine algal

Tetraselmis suecica attained a biomass production of 560 mg L-1 d-1 and a doubling time that

ranged from 26-28 h (Chini Zittelli et al., 2006). In experiments using the horizontal tubular

bioreactor for growth of Phaeodactylum tricornutum, biomass production was 1200 mg L-1 d-1

(Acién Fernandez et al., 2001) which doubles the biomass production of the bubble column

photobioreactor.

The horizontal tubular and bubble column photobioreactors provide different physical conditions

for growing algae and they include internal light and gas dispersion, mass transfer, surface

volume ratio, mixing, turbulence and nature of the turbulence. These factor are important

because they control availability of CO2 and light inside the reactor (Sanchez Miron et al., 1999;

Garcia Camacho et al., 1999).

Page 18: Effects of Light, Co2 and Reactor Design on Growth of Algae

 

The flat panel photobioeactor is another type of reactor system with chambers that are 2.4-10 cm

thick and a nutrient flow rate of 50 cm s-1. Its main advantage is less electrical energy

consumption than the tubular horizontal photobioreactor (Sierra et al., 2008). A modification of

the flat panel photobioreactor is the thin-layer reactor which has a reduced thickness to 0.6-1 cm

(Doucha and Livansky, 2009; Doucha et al., 2005; Livansky and Doucha, 1997). Table 1.1

summarizes photobioreactors systems.

Table 1.1. Algae biomass production in different growing systems.

System Algal specieBiomass Production   

(mg L‐1 day

‐1)

Source

Airlift tubular Porphyridium cruentum 1500 Camacho Rubio et al., 1999

Airlift tubular Phaeodactylum tricornutum 1200 Acién Fernández et al., 2001

Airlift tubular Phaeodactylum tricornutum 1900 Molina Grima et al., 2001

Inclined tubular Chlorella sorokiniana 1470 Ugwu et al., 2002

Undular row tubular Arthrospira platensis 2700 Carlozzi. 2003

Outdoor helical tubular Phaeodactylum tricornutum 1400 Hall et al., 2003

Parallel tubular (AGM) Haematococcus pluvialis 50 Olaizola. 2000

Bubble column Haematococcus pluvialis 60 Garcia‐Malea López et al., 2006

Flat plate Nannochloropsis sp 270 Cheng‐Wu et al., 2001

Semicontinuous system Nannochloropsis oculata 497 Chiu et al., 2009

Semicontinuous system Nannochloropsis oculata 296 Chiu et al., 2010

Outdoor photobioreactor Phaeodactylum tricornutum 300 Sánchez Mirón et al., 2003

Inclined tubular Chlorella sorokiniana 1470 Ugwu et al., 2002

Algal Biomass Production in different reactors

1.1.5 ILLUMINATION

1.1.5.1 FLASHING LIGHT AND DARK PERIOD

Photobioreactor can be illuminated either internally or externally from outside the reactor.

However, all illumination sources provide excessive light in areas close to the light source and

insufficient light for photosynthesis at distances away from the light source. Light dispersion is

not uniform (Molina et al., 1999) so that cells near to the light source receive more light than

Page 19: Effects of Light, Co2 and Reactor Design on Growth of Algae

 

those farther away (Frohlich et al., 1983). The cells themselves physically limit light dispersion,

producing shadows which block light transmission to other cells in the reactors and this reduces

photosynthesis (Ree and Gotham, 1981; Laws, 1980; Bannister, 1979; Tamiya et al., 1953). A

mathematical model that calculates the proportions of light vs dark zones in an algal reactor

culture concluded that the intensity of stirring does not significantly affect the proportion of non-

illuminated zones (Livasnsky, 1979). Therefore biomass production would not likely be

significantly enhanced by stirring under light limiting conditions

Since the reactions of light and dark photosynthesis occur continually, light can interfere with the

dark process of photosynthesis. Because of this, a more favourable growing condition is possible

when there is an alternation between periods of light and dark, and optimal light-dark

frequencies between 3 and 25 Hz have been reported (Perner-Nocha and Posten, 2007). Several

studies of altered light-dark frequencies have shown increased algal growth (Richmond et al.,

2003; Jannsen et al., 1999; Merchuck et al., 1998; Terry, 1986; Laws et al., 1983; Lee and Pirt,

1981).

Solar irradiation intensity can be as high as 2800 µmol m-2 s-1 which greatly exceeds the light

saturation point for algae photosynthesis of 50-300 µmol m-2 s-1 (Kim et al., 2006; Degen et al.,

2001; Merchuk et al., 1998). The flashing light technique has been suggested to improve light

utilization for outdoor algae culture (Terry, 1986). Under photoautotrophic growth the flashing

light frequency should be as high as possible for increased cell growth (Hu et al., 1996b).

Light utilization in flashes has been studied for algal photosynthesis with special emphasis on

flashing periods, the time interval between flashes, of less than 1 s and 100 ms (Terry, 1986;

Phillips and Myers, 1954). Flashing light periods at high frequencies (0.034 Hz or more) increase

the efficiency of the photosynthetic process (Sato et al., 2006). At these light flashing

frequencies and under optimal illumination intensities of 300 µmol m-2 s-1 to 1200 µmol m-2 s-1,

Sato et al. (2010) reported biomass production with Chaetoceros calcitrans and Chlorococum

Page 20: Effects of Light, Co2 and Reactor Design on Growth of Algae

 

littorale of 266 mg L-1 d -1 and 146 mg L-1 d -1, respectively. Flashing lights at frequencies of

0.001-200 Hz and light intensities of 12-18 µmol m-2 s-1 were used for Haematococcus pluvialis

production and the results show that astaxanthin concentrations were higher in cultures exposed

to flashing light frequencies higher than 1 Hz compared to algae growing under continuous light

(Katsuda et al., 2008). See Table 1.2 for studies conducted with algae growing under flashing

light treatments.

Table 1.2. Studies with algae under flashing lights.

Min Max

  Chlorella sp 0 750 0.3‐0.0074 1/1.0  Grobbelaar. 1991

  Scenedesmus sp 0 750 0.8‐0.0038 1/1.0  Grobbelaar. 1991

  Chlamydomonas reinhardtii 100 240 0.01 1.3/8.7  Janssen et al., 1999

  Chlorella sorokiniana 240 630 0.025 3/1.0   Janssen et al., 1999

  Haematococcus pluvialis 2 12 25‐200 1/1.0   Katsuda et al., 2006

  Phaeodactylum tricornutum 22 700 0.5‐1 1/10.0   Laws et al., 1983

  Chlorella pyrenoidosa 80 168 2.5‐25 kHz 1/10.0  Matthijs et al., 1996

  Phaeodactylum tricornutum 230 2800 1‐1.55 1/1.0  Molina Grima et al., 2000

  Chlorella kessleri 280 280 1‐37 kHz 1/1 to 1/20  Park and Lee. 2000

  Chlorella kessleri 50 78 10‐50 kHz 1/1 to 1/10   Park and Lee. 2001

  Phaeodactylum tricornutum 68 0.01‐10 1/1.0   Walsh and Legendre. 1988

  Chaetoceros calcitrans 10 188 0.1‐10  1/5 to 1/20   Yoshimoto et al., 2005

Flash Light and Light/Dark Period for Growing Algal Cells 

Algal specieIlumination  (µmol m

‐2S‐1)

Flash light (Hz) Light/Dark ratio Source

LED technology can provide light at a specific wavelength with 50-95% of the total light emitted

at that wavelength (Matthijs et al., 1996). Therefore, the use of LEDs for growing algae indoors

has increased due to their much lower energy consumption compared to fluorescent and

incandescent lamps. Matthijs et al. (1996) was able to grow Chlorella pyrenoidosa using red

LEDs, with a peak emission at 659 nm, and an alternating light period of 5 µs and dark period of

45 µs (a light/dark ratio of 1/10) under flashing conditions of 20 kHz.

There is evidence of quantum absorbing memory in algal photosynthesis. In an early study

conducted with Chlorella pyrenoidosa, a light flash for 1 ms at 61 µmol m-2 s-1 supported

photosynthesis for 20 ms (Philiph and Myers. 1954). More recently, studies have shown that the

Page 21: Effects of Light, Co2 and Reactor Design on Growth of Algae

 

energy stored by Chlorella vulgaris in a flash of 500 ms and at a light intensity of 58-73 µmol m-

2 s-1 was sufficient to support photosynthesis (assimilating CO2 and releasing O2) at maximum

rate for 9.2 ms (Lee and Pirt, 1981; Pirt, 1986). Table 1.3 shows a summary of optimal light

intensities for algal growth.

Table 1.3. Optimal light intensity for growing algae.

Min Max

  Chlorella pyrenoidosa 50 200   Ogbonna et al., 1999.

  Chlorella pyrenoidosa  (outdoor) 200 450  Ogbonna et al., 1999

  Euglena gracilis 50 450   Obbonna et al., 2001

  Chlorella vulgaris 10 250   Degen et al., 2001

  Chlorella vulgaris  (Outdoor) 80 980  Degen et al., 2001

  Chlamydomonas reinhardtii 60 200   Kim et al., 2006

  Chlamydomonas reinhardtii 200 300   Kim et al., 2006

  Phaeodactylum tricornutum 230 250   Fernández et al., 2003

  Phaeodactylum tricornutum  (outdoor) 800 1100  Fernández et al., 2003

  Porphyridium cruentum  (outdoor) 1600 2000  Rebolloso et al., 1999

  Haematococcus pluvialis 3.8 12   Lababpour et al., 2004

  Haematococcus pluvialis 3.8 12   Lababpour et al., 2005

  Haematococcus pluvialis 2 12   Katsuda et al., 2006

  Haematococcus pluvialis 21 49   Ranjbar et al., 2008

  Haematococcus pluvialis 23 60   Garcia‐Malea et al., 2005

  Haematococcus pluvialis  350 2500   Garcia‐Malea et al., 2005

  Haematococcus pluvialis  (Inside the culture) 70 130  Garcia‐Malea et al., 2006

  Haematococcus pluvialis  (outdoor) 50 2000  Garcia‐Malea et al., 2006

  Spirulina platensis 300 3000   Wang et al., 2007

  Nannochloropsis sp 100   Zittelli et al., 1999

Optimal Ilumination for Growing Algal Cells 

Algal specieIlumination  (µmol m

‐2S‐1)

Source

1.1.5.2 PHOTOINHIBITION IN ALGAE

High light emission intensities can cause photoinhibition, thereby reducing algal growth rate. In

a study conducted in the outdoors with Porphyridium cruentum it was observed that the

photosynthetic activity showed variation during the day period. The lowest photosynthetic

Page 22: Effects of Light, Co2 and Reactor Design on Growth of Algae

 

activity was observed between 12:00 and 2:00 pm, when solar radiation is usually highest

(Rebolloso Fuentes et al., 1999).

Studies of the biomass production of the marine microalgae Isochrysis galbana under light

intensities ranging from 820 to 3270 µmol m-2 s-1 showed a peak algal growth rate at 1630 µmol

m-2 s-1. Above this light intensity, biomass production decreased due to photoinhibition (Molina-

Grima et al., 1996). While biomass production of 19.2 mg L-1 h-1 at a light intensity of 3270

µmol m-2 s-1 was higher than that at 820 µmol m-2 s-1 (16.6 mg L-1 h-1), quantum yields decreased

above light intensities of 820 µmol m-2 s-1 (Molina-Grima et al., 1997). See Table 1.4 for

quantum yield absorption under different light intensities.

Table 1.4. Quantum yield in photon absorption by biomass under different photon flux.

Light intensity (µmol m-2 s-1)

Quantum yield (g biomass/energy mol )

820 0.603 1620 0.428 3270 0.087

Source: Molina-Grima et al., 1997)

Page 23: Effects of Light, Co2 and Reactor Design on Growth of Algae

 

1.1.5.3 WEIGHT LOSS DURING NIGHT PERIOD

Growing algae in controlled environments has problems related to the day and night cycles as

well as to variations of the light intensity during the day cycle. During night, photosynthetic cells

switch their biochemistry from energy production autotrophic growth to energy consumption

heterotrophic growth (Ogbonna et al., 1999). Consumption of this energy reserve during the

night has been reported to result in a cell body weight loss of 17% in Chlorella pyrenoidosa

(Ogbonna and Tanaka, 1996), and can represent up to 35% of the entire weight that is produced

during the day photosynthetic period (Torzillo et al., 1991a; Torzillo et al., 1991b; Grobbelarar

and Soeder, 1985). Experiments conducted with Spirulina platensis have shown that biomass

losses during the night period ranged from 5 to 10% and which represented losses of 16-32% of

the biomass daily production. These biomass losses are affected by growth temperature and light

irradiance during the day growth (Torzillo et al., 1991).

1.1.6 ALGAE GROWTH UNDER ARTIFICIAL AND OUTDOOR CONDITIONS

1.1.6.1 GROWTH UNDER INDOOR ARTIFICIAL CONDITIONS

Open ponds and photobioreactor is the most implemented technology to growth algae (Olivieri et

al., 2013). Algae can be grown under indoor artificial conditions using open ponds in

greenhouse, tubular bubble column, continous stirred tanks, horizontal tubular and flat panel

photobioreators. Sudies with the algal Chlorella vulgaris growing in tubular bubble column and

stirred tanks has produced standing biomass of 0.6-0.7 g L-1 (Sacasa-Castellanos, 2013). The

algae Spirulina platensis grown indoors in a horizontal tubular photobioreactor with a nutrient

media flow of 21 cm s-1 and illumination of 120 µmol m-2 s-1 produced a standing biomass of 10

g L-1 in 19 d. The same algal specie growing in an indoor open pond under similar conditions

attained a standing biomass of only 1.5 g L-1 in 20 d (Converti et al., 2006)

Page 24: Effects of Light, Co2 and Reactor Design on Growth of Algae

10 

 

1.1.6.2 GROWTH UNDER OUTDOOR CONDITIONS

Under outdoor conditions solar quantum yields during photosynthesis are less than 5%, and the

maximum biomass production is 30-40 g m-2 day-1 or 2-2.7 g L-1 day-1 (Goldman, 1979).

Experiments conducted during the summers 2003 and 2004 under outdoor conditions with the

algal Tetraselmis suecica attained biomass production of 0.49-0.56 g L-1 d-1 (Chini Zittelli et al.,

2006). Further experiments conducted with Nannochloropsis sp during the summer 1997 under

outdoor conditions reported biomass production of 0.51-0.76 g L-1 d-1 (Chini Zittelli et al., 1999).

Outdoor studies with the blue-green algae Anabaena variabilis (cyanobacteria) obtained biomass

production during winter of 0.40 g L-1 d-1 (6 g m-2 d-1) and 1.13 g L-1 d-1 (17 g m-2 d-1) in

summer, with photosynthetic activity occurring to 35 cm depth and temperatures controlled at

30-35oC (Fontes et al., 1987). Most recent outdoor experiments with Tetraselmis suecica and

Chlorella sp have obtained productivity of 0.18 and 0. 24 g L-1 d-1 respectivity (Moheimani,

2013).

1.1.7 COMMERCIAL PRODUCTION OF ALGAE

There has been commercial production of microalgae since 1950 in Taiwan and Japan

(Sivasubramanian et al., 2010; Muthukumaran et al., 2012). By 1997 there were aproximately

110 commercial enterprises of microalgae in the Asia Pacific region and the most common algae

produced were Chlorella sp, Nannochloris sp, Dunaliella sp, and Spirulina sp. (Lee, 1997). In

1996 the annual consumption of Chlorella sp in Japan was 2000 T of which 1057 T were

produced locally and 943 T were imported (Lee, 1997).

The most common uses of Chlorella sp biomass are as food supplements and in the comestic and

pharmaceutical industries (Sankar and Ramasubramanian, 2012); another significant use is oil

production for biofuels (Lyon et al., 2013). The top four commercial species of algal biomass are

3000 T for Spirulina sp, 2000 T for Chlorella sp, 1200 T for Dunnaliella salina, and 500 T for

Aphanizomenon sp (Priyadarshani and Rath, 2012). Another important species is the marine alga

Nannochloropsis salina, which has been considered for biodiesel production because its oil

production can exceed 50% of the body mass (Silva et al., 2014).

Page 25: Effects of Light, Co2 and Reactor Design on Growth of Algae

11 

 

Production of microalgal biodiesel has the potential to replace other sources of diesel fuel

without using soils destinated for agriculture (Rajvanshi and Sharma, 2012). The world΄s algae

production for biofuels can be divided in regions; 78% USA, 13% Europe and 9% in the rest of

the world. The most implemented technology is the closed bioreactor at 52% followed by open

ponds at 26%; only 22% of the production is grown under natural enviroments (Gendy and El-

temtamy, 2013). Some commercial companies that produce algae for biofuel in the USA are

Greenfuel Technologies Corporation, Solazyme, Livefuels, Solix Biofuels, and Inventure

Chemical (Gao et al., 2012).

1.1.8 ALGAE USED DURING THE COURSE OF THIS RESEARCH

This research examined the growth of three marine algae and three fresh water algae. Detailed

taxonomy of these algae is reported in Table 1.5.

Table 1.5. Taxonomy of the six algal species grown during the course of this research.

Classification Domain Eukaryota Eukaryota Eukaryota Eukaryota Eukaryota EukaryotaKingdom Protista Plantae Plantae Plantae Plantae PlantaeDivision Chlorophyta Chlorophyta Chlorophyta Chlorophyta Chlorophyta ChlorophytaClass Trebouxiophyceae Chlorophyceae Chlorophyceae Chlorodendrophyceae Chlorophyceae ChlorophyceaeOrder Chlorellales Sphaeropleales Chlamydomonadale Chlorodendrales Chlorococcales VolvocalesFamily Chlorellaceae Scenedesmaceae Chlamydomonadaceae Chlorodendraceae Coccomyxaceae DunaliellaceaeGenus Chlorella Scenedesmus Chlamydomonas Tetraselmis * Nannochloris * Dunaliella *Source: Arora et al., 2013; Assuncao et al., 2012; Huss et al., 1999; Krienitz et al., 1996.

Fresh water algae Marine algae *Classification of the algae studied in this research

The algae used during this research have different size and shape, Figure 1.1 shows the features

of these algae.

Page 26: Effects of Light, Co2 and Reactor Design on Growth of Algae

12 

 

Figure 1.1. Features of the six algal species studied during the course of this research.

1.2 FORMAT OF THE THESIS

The thesis is contained in 10 chapters: Chapter one is the introduction for the thesis and chapter

two states the objectives and goal of the research. Chapter three contains the literature review

including the evolution of the photosynthetic machinery anoxygenic, oxygenic, different types of

photosynthetic organisms, as well as a detailed description of the architecture and function of

PSII and PSI. Chapter four describes the growth of two marine and two fresh water algae under

solar radiation, white light, blue LED light and red/blue LED light and the combination of the

treatments to determine the optimal light treatment to support algae growth. Chapter five

discusses further research in light treatments versus biomass production that combines 12 hours

solar radiation against 24 hours red/blue LED and 12 hours solar plus 12 hours red/blue LED in

six different algae species. Chapter six describes a study with Chlorella vulgaris under four

different 24 hours LED light treatments; white, red, blue and red/blue we measured population

density and biomass production. Chapter seven provides an artificial illumination experiment

using light and dark periods in milliseconds: the flashing effect was used to growth one marine

and one fresh water algae against the same algae under continuous red/blue LED. Chapter eight

Page 27: Effects of Light, Co2 and Reactor Design on Growth of Algae

13 

 

consists in a group of experiments that measures different CO2 concentration in the growth and

biomass production of Chlorella vulgaris, the goal of the experiment is to determine if CO2

concentration in the air supply affects algae biomass production. Chapter nine describes the

study of two marine and two fresh water algae under a flat panel reactor using natural light

during day periods and artificial light during night periods, these study also measured the effect

of CO2 concentration and biomass production in an industrial scale. Chapter ten presents the

overall conclusions of all six different studies in algae.

Page 28: Effects of Light, Co2 and Reactor Design on Growth of Algae

14 

 

1.3 REFERENCES

1. Acien Fernandez, F. G., Fernandez Sevilla, J. M., Sanchez Perez, J. A., Molina-Grima, E., Chisti, Y. 2001. Airlift-driven external-loop tubular photobioreactors for outdoor production of microalgae: assessment of design and performance. Chem. Eng. Scien. 56: 2721-2732.

2. Arora, M., Anil, A.C., Leliaert, F., Delany, J., Mesbahi, E. 2013. Tetraselmis indica

(Chlorodendeophyceae, Chlorophyta), a new species isolated from salt pans in Goa, India. Eur. J. Phycol. 48: 61-78.

3. Assuncao, P., Jaen-Molina, R., Caujape-Castells, J., de la Jara, A., Carmona, L.,

Freijanes, K., Mendoza, H. 2012. Molecular taxonomy of Dunaliella (Cholorophyceae), with a special focus on D. Salina: ITS2 sequences revisited with an extensive geographical sampling. Aquat. Biosyst. 8-2: 1-13.

4. Bannister, T.T. 1979. Quantitative description of steady state, nutrient-saturated algal growth, including adaptation. Limnol. Ocean. 24: 76-96.

5. Baquerisse, D., Nouals, S., Isambert, A., Ferreira Dos Santos, P., Durand, G. 1999. Modelling of a continuous pilot photobioreactor for microalgae production. J. Biotechnol. 70: 335-342.

6. Sacasa-Castellanos, C. 2013. Batch and continuous studies of Chlorella vulgaris in

photo-bioreactors. The University of Western Ontario, 67p.

7. Chini Zittelli, G., Rodolfi, L., Biondi, N., Tredici, M. R. 2006. Productivity and photosynthetic efficiency of outdoor cultures of Tetraselmis suecica in annular columns. Aquaculture. 261: 932-943.

8. Converti, A., Lodi, A., Del Borghi, A., Solisio, C. 2006. Cultivation of Spirulina platensis in a combined airlift-tubular reactor system. Biochem. Eng. J. 32: 13-18.

Page 29: Effects of Light, Co2 and Reactor Design on Growth of Algae

15 

 

9. Degen, J., Uebele, A., Retze, A., Schmid-Staiger, U., Trosch, W. 2001. A novel airlift photobioreactor with baffles for improved light utilization through the flashing light effect. J. Biotechnol. 92: 89-94.

10. Doucha, J., Livansky, K. 2009. Outdoor open thing-layer microalgal photobioreactor: potential productivity. J. Appl. Phycol. 21: 111-117.

11. Doucha, J., Straka, F., Livansky, K. 2005. Utilization of flue gas for cultivation of microalgae (Chlorella sp) in an outdoor open thin-layer photobioreactor. J. Appl. Phycol. 17: 403-412.

12. Fan, L. H., Zhang, Y. T., Zhang, L., Chen, H. L. 2008. Evaluation of a membrane-sparged helical tubular photobioreactor for carbon dioxide biofixation by Chlorella vulgaris. J. Membrane Scien. 325: 336-345.

13. Fenical, W. 1997. New pharmaceuticals from marine organisms. Tibtech. 15: 339-341.

14. Fernandez, F. G., Hall, D. O., Guerrero, E., Rao, K. K., Molina-Grima, E. 2003. Outdoor production of Phaeodactylum tricornutum biomass in a helical reactor. J. Biotechnol. 103-2: 137-152.

15. Fontes, A. G., Angeles Vargas, M; Moreno, J., Guerrero, M. G., Losada, M. 1987. Factor affecting the production of biomass by a nitrogen-fixing blue-green alga in outdoor culture. Biomass. 13: 33-43.

16. Frohlich, B.T., Webster, I.A., Ataai, M.M., Shuler, M.L. 1983. Photobioreactors: models for interaction of light intensity, reactor design and algal physiology. Biotechnol. Bioeng. 13: 331-350.

17. Gao, Y., Gregor, C., Liang, Y., Tang, D., Tweed, C. 2012. Algae biodiesel-a feasibility report. Chem. Cent. J. 6: 1-16.

Page 30: Effects of Light, Co2 and Reactor Design on Growth of Algae

16 

 

18. Garcia Camacho, F., Contreras Gomez, A., Acien Fernandez, J., Fernandez Sevilla, J., Molina Grima, E. 1999. Use of concentric-tube airlift photobioreactors for microalgal outdoor mass cultures. Enzyme Microb. Tech. 24: 164-172.

19. Garcia-Malea Lopez, M.C., Del Rio Sanchez, E., Casas Lopez, J,L., Acien Fernandez, F.G., Fernandez Sevilla, J.M., Rivas, J., Guerrero, M.G., Molina Grima, E. 2006. Comparative analysis of the outdoor culture of Haematococcus pluvialis in tubular and bubble column photobioreactors. J. Biotechnol. 123: 329-342.

20. García-Malea, M.C., Brindley, C., Del Rio, E., Acién, E. G., Fernández, J.M., Molina, E. 2005. Modeling of growth and accumulation of carotenoids in Haematococcus pluvialis as a function of irradiance and nutrient supply. Biochem. Eng. J. 26: 107-114.

21. Gendy, T.S., El-Temtamy, S.A. 2013. Comercialization potential aspects of microalgae for biofuel production: an overview. Egypt. J. Pet. 22: 43-51.

22. Goldman, J. C. 1979. Outdoor algal mass cultures: II photosynthetic yield limitations. Water Res. 13: 119-136.

23. Grobbelaar, J.U. 1991. The influence of light/dark cycles in mixed algal cultures on their productivity. Bioresour. Technol. 38: 189-194.

24. Grobbelaar, J.U., Soeder, C.J. 1985. Respiration losses in green alga cultivated in raceway ponds. J. Plankton Res. 7: 497-506.

25. Hu, Q., Guterman, H., Richmon, A. 1996(A). A flat inclined modular photobioreactor for outdoor mass cultivation of photoautotrophs. Biotechnol. Bioeng. 51: 51-60.

26. Hu, Q., Richmond, A. 1996(B). Productivity and photosynthetic efficiency of Spirulina platensis affected by light intensity, cell density and rate of mixing in flat plate photobioreactor. J. Appl. Physiol. 8: 139-145.

Page 31: Effects of Light, Co2 and Reactor Design on Growth of Algae

17 

 

27. Huss, V.A.R., Frank, C., Hartmann, E.C., Hirmer, M., Kloboucek, A., Seidel, B.M., Wenzeler, P., Kessler, E. 1999. Biochemical tazonomy and molecular phylogeny of the genus Chlorella sensu lato (Chlorophyta). J. Phycol. 35: 587-598.

28. Issarapayup, K., Powtongsook, S., Pavasant, P. 2009. Flat panel airlift photobioreactors for cultivation of vegetative cells of microalga Haematococcus pluvialis. J. Biotechnol. 142: 227-232.

29. Jannsen,M., Kuijpers, T., Veldhoen, B., Ternbach, M., Tramper, J., Mur, L., Wijffels, R. 1999. Specific growth rate of Chlamydomonas reinhardtii and Chlorella sorokiniana under medium duration light/dark cycles. J. Biotechnol. 70: 323-333.

30. Katsuda, T., Shimahara, K., Shiraishi, H., Yamagami, K., Ranjbar, R., Katoh, S. 2006. Effect of flashing light from blue light emitting diodes on cell growth and astaxanthin production of Haematococcus pluvialis. J. Bioscien. Bioeng. 102-5: 442-446.

31. Katsuda, T., Shiraishi, H., Ishizu, N., Ranjbar, R., Katoh, S. 2008. Effect of light intensity and frequency of flashing light from blue light emitting diodes on astaxanthin production by Haematococcus pluvialis. J. Bioscien. Bioeng. 105-3: 216-220.

32. Kim, J.P., Kang, C.D., Park, T.H., Kim, M.S., Sim, S.J. 2006. Enhanced hydrogen production by controlling light intensity in sulphur-deprived Chlamydomonas reinhardtii culture. Int. J. Hydrogen Energy. 31: 1585-1590.

33. Krienitz, L., Huss, V.A.R., Hummer, C. 1996. Picoplanktonic Choricystis species

(Chlorococcales, Chlorophyta) and problems surrounding the morphologically similar Nannochloris-like algae. Phycologia. 35-4: 332-341.

34. Lababpour, A., Shimashara, K., Hada, K., Kyoui, Y., Katsuda, T., Katoh, S. 2005. Fed-batch culture under illumination with blue light emitting diodes (LEDs) for astaxanthin production by Haematococcus pluvialis. J. Bioscien. Bioeng. 100-3: 339-342.

Page 32: Effects of Light, Co2 and Reactor Design on Growth of Algae

18 

 

35. Lababpour, A., Hada, K., Shimahara, K., Shimahara, K., Katsuda, T., Katoh, S. 2004. Effects of nutrients supply methods and illumination with blue light emitting diodes (LEDs) on astaxanthin production by Haematococcus pluvialis. J. Bioscien. Bioeng. 98-6: 452-456.

36. Laws, E.A., Terry, K.L., Wickman, J., Chalup, M.S. 1983. A simple algal production system designed to utilize the flashing light effect. Biotechnol. Bioeng. 25: 2319-2335.

37. Laws, E.A. 1980. Nutrient and light limited growth of Thalassiosira fluviatilis in continuous culture with implications for phytoplankton growth in the ocean. Limnol. Ocean. 25: 455-473.

38. Lee, H.S., Seo, M. W., Kim, Z. H., Lee, C. G. 2006. Determining the best specific light uptake rates for the lumostatic cultures in bubble column photobioreactors. Enzyme Microb. Tech. 39: 447-452.

39. Lee, Y. 1997. Comercial production of microalgae in the Asia-Pacific rim. J. Appl. Phyc. 9: 403-411.

40. Lee, Y., Pirt, S.J. 1981. Energetics of photosynthetic algal growth: influence of intermittent illumination in short (40 s) cycles. J. Gen. Microbiol. 124: 43-52.

41. Livansky, K., Doucha, J. 1997. Additional CO2 saturation of thin-layer outdoor microalgal cultures: CO2 mass transfer and absorption efficiency. Algol Stud. 87: 145-154.

42. Livansky, K. 1979. Effect of the nonilluminated part of suspension on biomass production in an algal reactor. Folia Microbiol. 24: 339-345.

43. Lucumi, A., Posten, C. 2006. Establishment of long-term perfusion cultures of recombinant moss in a pilot tubular photobioreactor. Process Biochem. 41: 2180-2187.

Page 33: Effects of Light, Co2 and Reactor Design on Growth of Algae

19 

 

44. Lyon, T.A., Primm, J.G., Wojdan, M.S., Morehouse, K.D., Grogger, M.L., Vintila, S., Veverka, D.V. 2013. Environmental manipulation of select algae strains for maximal oil production. Bios. 84-1: 21-29.

45. Matthijs, H. C. P., Balke, H., Van Hes, U. M., Kroon, B. M. A., Mur, L. R., Binot, R. 1996. Application of light-emitting diodes in bioreactors: flashing light effects and energy economy in algal culture Chlorella pyrenoidosa. Biotechnol. Bioeng. 50: 98-107.

46. Meireles, L. A., Guedes, A. C., Barbosa, C. R., Azevedo, J. L., Cunha, J. P., Malcata, F. X. 2008. On-line control of light intensity in a microalgal bioreactor using a novel automatic system. Enzyme Microb. Tech. 42: 554-559.

47. Merchuk, J.C., Ronen, M., Giris, S., Arad, S. 1998. Light/dark cycles in the growth of the microalga Porhyridium sp. Biotechnol. Bioeng. 59: 706-713.

48. Moheimani, N. R. 2013. Long-term outdoor growth and lipid productivity of Tetraselmis suecica, Dunaliella tertiolecta and Chlorella sp (Chlorophyta) in bag photobioreactor. J. Appl. Phycol. 25: 167-176.

49. Molina Grima, E., Acien Fernandez, F.G., Garcia Camacho, F., Camacho Rubio, F., Chisti, Y. 2000. Scale-up of tubular photobioreactors. Journal of Appl. Phycol. 12: 355-368.

50. Molina Grima, E., Acien Fernandez, F.G., Garcia Camacho, F., Chisti, Y. 1999. Photobioreactors: light regime, mass transfer, and scaleup. J. Biotechnol. 70: 231-247.

51. Molina Grima, E., Garcia Camacho, F., Sanchez Perez, J.A., Acien Fernandes, F.G., Fernandez Sevilla, J.M. 1997. Evaluation of photosynthetic efficiency in microalgal cultures using averaged irradiance. Enzyme Microb. Tech. 21: 375-381.

52. Molina-Grima, E., Fernandez-Sevilla, J.M., Sanchez-Perez, J.A., Garcia-Camacho, F. 1996. A study of simultaneous photolimitation and photoinhibition in dense microalgal cultures taking into account incident and averaged irradiances. J. Biotechnol. 45: 59-69.

Page 34: Effects of Light, Co2 and Reactor Design on Growth of Algae

20 

 

53. Muthukumaran, M., Thirupathi, P., Chinnu, K., Sivasubramanian, V. 2012. Phycoremediation efficiency and biomass production by micro alga Desmococcuc olivaceus (Persoon et Acharius) J.R. Laundon treated on chrome-sludge from an electroplating industry-A open raceway pond study. Int. J. Curr. Sci. 1: 52-62.

54. Ogbonna, J. C., Soejima, T., Ugwu, C. U., Tanaka, H. 2001. An integrated system of solar light, artificial and organic carbon supply for cyclic photoautotrophic-heterotrophic cultivation of photosynthetic cells under day-night cycles. Biotechnol. Lett. 23: 1401-1406.

55. Ogbonna, J.C., Soejima, T., Tanaka, H. 1999. An integrated solar and artificial light system for internal illumination of photobioreactors. J. Biotechnol. 70: 289-297.

56. Ogbonna, J.C., Tanaka, H., 1996. Night biomass loss and changes in biochemical composition of cells during light: dark cyclic culture of Chlorella pyrenoidosa. J. Ferm. Bioeng. 82: 558–564.

57. Olivieri, G., Salatino, P., Marzocchella, A. 2013. Advances in photobioreactors for intensive microalgal production: configurations, operating strategies and applications. J. Chem. Technol. Biotechnol. X: x-x.

58. Park. K.H., Lee, C.G. 2000. Optimization of algal photobioreactors using flashing lights. Biotechnol. Bioprocess Eng. 5: 186-190.

59. Perner-Nocha, I., Clemens, P. 2007. Simulations of light intensity variation in photobioreactors. J. Biotechnol. 131: 276-285.

60. Phillips, J.N., Myers, J. 1954. Growth rate of Clorella in flashing light. Plant Physiol. 29: 152-161.

61. Pirt, S.J. 1986. The thermodynamic efficiency (quantum demand) and dynamics of photosynthetic growth. New Phytol. 102: 3-37.

Page 35: Effects of Light, Co2 and Reactor Design on Growth of Algae

21 

 

62. Priyadarshani, I., Rath, B. 2012. Commercial and industrial application of micro algae – a review. J. Algal Biomass Utln. 3-4: 89-100.

63. Rajvanshi, S., Sharma, M.P. 2012. Microalgae: A potential source of biodiesel. J. Sustainable Bioenergy Syst. 2: 49-59.

64. Ranjbar, R., Inoue, R., Shiraishi, H., Katsuda, T., Katoh, S. 2008. High efficiency production of astaxanthin by autotrophic cultivation of Haematococcus pluvialis in a bubble column photobioreactor. Biochem. Eng. J. 39-3: 575-580.

65. Rebolloso Fuentes, M.M., Garcia Sanchez, J.L., Fernandez Sevilla, J. M., Fernandez Acien, F. G., Sanchez Perez, J. A., Molina Grima, E. 1999. Outdoor continuous culture of Porphyridium cruentum in a tubular photobioreactor: quantitative analysis of the daily cyclic variation of culture parameters. J. Biotechnol. 70: 271-288.

66. Ree, G.Y., Gotham, I.J. 1981. The effect of environmental factors on phytoplankton growth: Light and interaction of light with nitrate limitation. Limnol. Ocean. 26: 649-659.

67. Richmond, A., Cheng-Wu, Z., Zarmi, Y. 2003. Efficient use of strong light for high photosynthetic productivity: interrelationships between the optical path, the optimal population density and cell-growth inhibition. Biomol. Eng. 20: 229-236.

68. Rocha, J. M. S., Garcia, J. E. C., Henriques, M. H. F. 2003. Growth aspects of the marine microalga Nannochloropsis gaditana. Biomol. Eng. 20: 237-242.

69. Sankar, M., Ramasubramanian, V. 2012. Biomass production of commercial algae Chlorella vulgaris on different culture media. ISSN. 1-1: 56-60.

70. Sato, T., Yamada, D., Hirabayashi, S. 2010. Development of a virtual photobioreactor for microalgae culture considering turbulent flow and flashing light effect. Energy Convers. Manage. 51: 1196-1201.

Page 36: Effects of Light, Co2 and Reactor Design on Growth of Algae

22 

 

71. Sato,T., Usui, S., Tsuchiya, Y., Kondo, Y. 2006. Invention of outdoor closed type photobioreactor for microalgae. Energy Convers. Managet. 47: 791-799.

72. Sanchez Miron, A., Contreras Gomez, A., Garcia Camacho, F., Molina Grima, E., Chisti, Y. 1999. Comparative evaluation of compact photobioreactors for large-scale monoculture of microalgae. J. Biotechnol.70: 249-270.

73. Scragg, A. H., Illman, A. M., Carden, A., Shales, S. W. 2002. Growth of microalgae with increased calorific values in a tubular bioreactor. Biomass Bioenerg. 23: 67-73.

74. Sierra, E., Acien, F.G., Fernandez, J.M., Garcia, J.L., Gonzales, C., Molina, E. 2008. Characterization of a flat photobioreactor for the production of microalgae. Chem. Eng. J. 138: 136-147.

75. Silva, C., Soliman, E., Cameron, G., Fabiano, L.A., Seider, W.D., Dunlop, E.H., Coaldrake, A.K. 2014. Commercial-Scale biodiesel production from algae. Ind. Eng. Chem. Res. 53: 5311-5324.

76. Sivasubramanian, V., Subramanian, V.V., Muthukumaran. 2010. Bioremedation of chrome-sludge from an electroplating industry using the micro alga Desmococcus olivaceus a pilot study. J. Algal Biomass Utln. 1-3: 104-128.

77. Tamiya, H., Hase, E., Shibata, K., Mituya, A., Iwamura, T., Nihei, T., Sasa, T. 1953. Kinetics of growth of Chlorella, with special reference to its dependence on quantity of available light and on temperature. Algal culture from laboratory to pilot plant. Carnegie Institution of Washington, Washington, DC. 234-232.

78. Terry, K. 1986. Photosynthesis in modulated light: quantitative dependence of photosynthetic enhancement on flashing rate. Biotechnol. Bioeng. 28: 988-995.

79. Torzillo, G., Sacchi, A., Materassi, R. 1991a. Temperature as an important factor affecting productivity and night biomass loss in Spirulina platensis grown outdoors in tubular photobioreactors. Bioresour. Technol. 38: 95-100.

Page 37: Effects of Light, Co2 and Reactor Design on Growth of Algae

23 

 

80. Torzillo, G., Sacchi, A., Materassi, R., Richmond, A. 1991b. Effect of temperature on yield and night biomass loss in Spirulina platensis grown outdoor in tubular photobioreactors. J. Appl. Physiol. 3: 103-109.

81. Walsh, P., Legendre, L. 1988. Photosynthetic responses of the diatom Phaeodactylum tricornutum to high frequency light fluctuations simulating those induced by sea surface waves. J. Plankton Res. 10-5: 1077-1082.

82. Wang, C.Y., Fu, C.C., Liu, Y.C. 2007. Effects of using light-emitting diodes on the cultivation of Spirulina platensis. Biochem. Eng. J. 37: 21-25.

83. Watanabe, Y., Saiki, H. 1997. Development of a photobioreactor incorporating Chlorella sp. for removal of CO2 in stack gas. Energy Convers. Manage. 38: S499-S503.

84. Wright, P. C., Stevenson, C., Mcevoy, E., Burgess, G. J. 1999. Opportunities for marine bioprocess intensification using novel bioreactor design: frequency of barotolerance in microorganisms obtained from surface waters. J. Biotechnol. 70: 343-349.

85. Yoshimoto, N., Sato, T., Kondo, Y. 2005. Dynamic discrete model of flashing light effect in photosynthesis of microalgae. J. Appl. Phycol. 17: 207-214.

86. Zittelli, G., Lavista, F., Bastianini, A., Rodolfi, L., Vincenzini, M., Tredici, M.R. 1999. Production of eicosapentaenoic acid by Nannochloropsis sp. cultures in outdoor tubular photobioreactors. J. Biotechnol. 70: 299–312.

Page 38: Effects of Light, Co2 and Reactor Design on Growth of Algae

24 

 

 

CHAPTER 2 RESEARCH OBJECTIVES

2.1 GENERAL GOAL

The goal of this thesis research is to study the critical factors affecting the growth of marine and

fresh water microalgae in a bench-top and pilot-scale. Factors studied were light source and

intensity, CO2 supply concentration and bioreactor design. Algal growth was accessed by

measurements of growth rate, biomass production, cell size and cell weight. We also constructed

and tested a large-scale flat panel photo-bioreactor that overcomes some of the biomass

production limitations common to bubble column photo-bioreactors.

2.2 SPECIFIC RESEARCH OBJECTIVES

1. To measure growth of marine and fresh water algae under different light sources: solar,

white, red, blue and the combination red-blue.

2. To evaluate the flashing light effects on the growth and biomass production of marine

and fresh water algae.

3. To measure the effects of increasing CO2 concentrations on algal growth kinetics and

biomass production.

4. To design and test a flat panel photo-bioreactor for biomass synthesis using marine and

fresh water algae.

Page 39: Effects of Light, Co2 and Reactor Design on Growth of Algae

25 

 

 

CHAPTER 3 THE PHOTOSYNTHETIC MACHINERY: A DETAILED

DESCRIPTION OF THE LIGHT REACTION; PSI, PSII, ATP SYNTHASE, STATE OF TRANSITION, PHOTO-DAMAGE, REGENERATION, AND MACHINERY ARCHITECTURE

3.1 ABSTRACT

The overall photosynthetic efficiency from solar radiation is in the range of 4-6%, this

electromagnetic energy hits and energizes the antenna of photosynthetic organisms. The antenna

pass the energy as vibrational energy from chlorophyll to chlorophyll until reaches the reaction

center (RC), inside the reaction center the special pair of chlorophyll receive this energy and

send it into an electron that passes from a ground state to an excite state. This electron leaves the

reaction center using the mobiles carrier; via plastoquinone b, plastocyanin and cytochrome b6f

in PSII. In PSI the route is via phylloquinone, ferredoxin, ferredoxin-NADP-reductase known as

FNRI. In anoxygenic photosynthesis in the purple bacteria the route is from the reaction center

PSII passing through Bacterio-Pheophytin-plastoquinone Qa to cytochrome bc1. The evolution

of photosynthesis has changed from one that was anoxygenic and energized the stage PSII

complex with a reverse electron transfer to produce NADPH, as in purple sulfur and non-sulfur

bacteria, to only one energized state in green sulfur bacteria ending with oxygenic photosynthesis

with two energize stages consisting in PSII and PSI, as found in cyanobacteria, algae and all

plants including single cell and multicellular structures.

GLOSSARY

Anoxygenic photosynthesis: Photosynthesis that is performed in the absence of oxygen. The process requires light, a source of carbon (CO2, acetate, lactate or pyruvate), an electron donor (H2 or H2S). In some cases water is produced as a byproduct. It is performed by bacteria and heliobacteria.

Page 40: Effects of Light, Co2 and Reactor Design on Growth of Algae

26 

 

Rhodopsin photosynthesis: Photosynthesis performed under aerobic or anaerobic conditions; light absorption is driven by the protein called rhodopsin. It is performed by Archaea and Bacteria.

Oxygenic photosynthesis: Photosynthesis in the presence of oxygen. The process requires light, a source of carbon (CO2), an electron donor (H2O), and produces oxygen as byproduct. It is performed by cyanobacteria, algae and plants.

Photosystem I (PSI): Secondary machinery that energized the electron for the second time during oxygenic photosynthesis to produce NADPH.

Photosystem II (PSII): Primary machinery in oxygenic photosynthesis that splits the water molecule and energizes the electron from water to produce ATP.

Light harvesting complex (LHCI, LHCII, or Antennae): Structure that absorbs electromagnetic energy, converts this energy into vibrational energy and transfers the energy to the reaction center (RC).

Reaction center (RC): Protein structure where the electron is energized during photosynthesis for the production of ATP or NADPH.

Water oxidation complex (WOC) and Oxygen evolution center (OEC): structures where water is split and oxygen produce during oxygenic photosynthesis in PSII.

D1 and D2 proteins: Major proteins of PSII where the RC, pheophythin D1-D2, WOC, OEC, and other important structures are located including the plastoquinones Qa and Qb (mobile carriers that pass the energized electron to cytochrome b6f).

PSaA and PSaB: Major proteins of PSI where the RC, special pair of chlorophylls, Chlorophyll A0a, A0b and other structures are located.

Light independent reaction (Calvin cycle): In oxygenic photosynthesis this process fixes CO2 and uses the energy from ATP and NADPH to synthetize G3P (Glyceraldehyde 3 phosphate), that is the precursor of glucose.

3.2 INTRODUCTION

The amount of solar radiation entering the earth’s atmosphere is ±1376 W m-2 and at the earth’s

surface on the equator is ± 1000 W m-2 (Kruse et al., 2005). However, the total energy available

for the photosynthetic activity of plants, cyanobacteria and algae is only 43% of this (430 W m-

2), with 5% in the ultraviolet region and 52% in the infrared region. Part of the energy in the

infrared region can be used in anoxygenic photosynthesis so that purple and green bacteria using

Page 41: Effects of Light, Co2 and Reactor Design on Growth of Algae

27 

 

440 W m-2 (Kruse et al., 2005). The efficiency of intercepted light by the photosynthetic

pigments ranges from 6 to 11%, the overall efficiency considering 76% of interception is 4-8%

(Wilson et al., 1991). Other estimates of the net photosynthetic efficiency are in the range of 4-

6% (Zhu et al., 2010; Zhu et al., 2008). Mejia (2006) calculated an overall photosynthetic

efficiency of 6.6% with an energy losses of 93.4%.

In photo-autotrophs, photosynthesis can operate in all earth environments where solar radiation

is present, in oceans, lakes, land and soil, and in aerobic or anaerobic environments. The

photosynthetic organisms are divided into oxygenic, anoxygenic and rhodopsin belonging the the

archaea and bacteria domain (Staley, et al 2007). The most common photosynthetic organisms

are aerobic which carried out oxygenic photosynthesis. In general photosynthesis captures light

energy and converts this energy to oxidise water, reduce ADP to ATP and NADPH synthesis, as

well as for O2 production by splitting water. The entire system is based on the necessity for an

inorganic carbon source, CO2. The ATP and NADPH are used to synthetize G3P in the light

independent reaction (Ferreira et al., 2004).

Oxygenic photosynthesis is performed by eukaryotic organisms such as plants, multicellular

algae, single cell algae and bacteria like cyanobacteria. In oxygenic photosynthesis H2O is used

as an electron donor, CO2 and the H2 from the split of H2O are used as electron acceptors in the

formation of glyceraldehyde-3-phosphate (G3P) during the Calvin cycle. This process also

produces O2 during the splitting of H2O in the light reactions, and therefore is called oxygenic

photosynthesis. (Staley et al., 2007; Brock 1979).

Anoxygenic photosynthesis is carried out in a non-oxygen environment by a group of

microorganisms called purple non-sulfur bacteria, green and purple sulfur bacteria, and by the

heliobacteria. In anoxygenic photosynthesis H2 or H2S are the electron donors; they are oxidised

and CO2 is reduced during the Calvin cycle to form G3P, and H2O is produced as a by-product

instead of O2. Anoxygenic photosynthesis by heliobacteria (bacteriochlorophyll g 788 nm) does

not use CO2 as carbon sources but uses monomers such as acetate, lactate and pyruvate. These

bacteria grow in soils and do not grow photoautotrophically. There is also a group of organisms

belonging to the archaea and bacteria domains that perform photosynthesis in aerobic and

anaerobic environments. (Staley et al., 2007; Brock 1979).

Page 42: Effects of Light, Co2 and Reactor Design on Growth of Algae

28 

 

Rhodopsin photosynthesis do not uses carotenoids neither chlorophyll nor bacterio-chlorophyll,

to absorb light for photosynthesis. The photosynthetic machinery have Bacteriorhodopsin with

absorption peak of 570 nm. These types of photosynthetic organisms do not have an electron

transport chain to produce NADPH during the light reaction, but use bacteriorhodopsin

molecules to create the proton gradient necessary to run the ATP synthase which makes ATP

(Staley et al., 2007; Brock 1979).

3.3 EVOLUTION OF PHOTOSYNTHESIS: ANOXYGENIC AND OXYGENIC

The first photosynthetic organisms on earth appeared 3-3.5 billion years ago, and were

anoxygenic organisms from the bacteria and archaea domains. These microorganisms included

the purple and green sulfur and nonsulfur bacteria and those possessing rhodopsin

photosynthesis, the organisms inhabited the anaerobic environment present on earth at that time.

The first oxygenic organisms to evolve were from the cyanobacteria group, and appeared 2.6-2.8

billion years ago during the Proterozoic period (Staley et al., 2007). These cyanobacteria had the

ability to survive in an anaerobic environment even though they produced O2. All of the O2

produced at that time oxidised Fe2+ to Fe2O3, a process that took place over 0.5 to 1 billion years.

Oxygen began to accumulate and the ozone layer to appear in the atmosphere 2 billion years ago.

There was an evolution among the purple sulfur bacteria and the green sulfur bacteria; the purple

bacteria used a reverse electron transfer reaction to capture sufficient energy to synthetize

NADPH; the green sulfur bacteria had a bacteriochlorophyll that allowed them to harvest energy

to produce ATP and NADPH without using the reverse electron transfer reaction (Staley et al.,

2007).

The photosynthetic machinery of the anoxygenic photosynthesis bacteria is much less complex

than that of oxygenic photosynthesis. For example, purple bacteria has the same mechanisms to

produce ATP as cyanobacteria with different photosynthetic pigments, however the purple sulfur

and nonsulfur bacteria use the reverse electron transfer reaction to synthetize NADPH. The green

sulfur bacteria and cyanobacteria do not need to use this mechanism because both have evolved

to harvest sufficient energy to produce ATP and NADPH. Cyanobacteria use a two step electron

Page 43: Effects of Light, Co2 and Reactor Design on Growth of Algae

29 

 

excitation, known as photosynthesis II (P 680) to produce the proton gradient for ATP synthesis,

and photosynthesis I (P 700) to produce the remaining energy needed to synthetize NADPH; this

represents an evolution from the purple bacteria (Staley et al., 2007).

The Eukarya domain appeared 600 million of years ago, and the photosynthetic organism that we

know today (single cell algae, multicellular algae and plants) evolved from a symbiosis of

eukaryotic primitive cells and cyanobacteria. It has been hypothized that cyanobacteria enter to

the cell and transformed into what we know today as chloroplast. Evidence from this symbiotic

union comes from the 16S rRNA analysis which shows that chloroplast has their own DNA

different from the eukaryotic cell’s nucleus. The chloroplast ribosome DNA is 70s which is a

similar type of DNA found in cyanobacteria and all bacteria domain, besides chloroplast

organelles have DNA with no nucleus envelope like prokaryotic cells. Another evidence is that a

typical ribosome DNA from 16S rRNA in eukaryotic cells is 80S. In chloroplast the internal

membranes containing the chlorophyll molecules, is called chlorosomes and they are found in

algae and in plants cells with another name thylakoid disk. All this evidence show that

chloroplast are an evolution of the cyanobacteria group and primitive eucaryotic cells.

Chloroplast uses the same chlorophyll found in cyanobacteria; Chlorophyll a antenna and

chlorophyll a reaction center to gather energy to excite an electron that goes in an electron

transport chain to produce ATP and NADPH, this system is similar in cyanobacteria, algae and

plants. The only difference is the size of the cyanobacteria cells which are much smaller than

plants and algae cells. These evidence lead to the hypothesis that chloroplast one day were free

living cyanobacteria and the symbiosis between eukaryotic organisms with cyanobacteria allow

eukaryotic cells to have a photosynthetic mechanisms to produce energy from inorganic

compounds. This explain why cyanobacteria and all the photosynthetic eukaryots perfomer the

same photosynthetic mechanisms pathways (Staley et al 2007).

There is a special pair of chlorophylls in the reaction center in most of the photosynthetic

organisms, the function of the reaction center is to charge an electron using the energy from the

antenna, and to donate this excited-state electron to the pheophytin protein to drive the

production of ATP and NADPH (Barber and Archer, 2001). The evolution of photosynthesis can

be traced with the boost in the oxidation potential of the reaction center. The special pair of

Page 44: Effects of Light, Co2 and Reactor Design on Growth of Algae

30 

 

chlorophyll in the reaction center of anoxygenic photosynthesis in green sulfur bacteria which is

binding P840 has an oxidation potential of 0.4 V and in purple non sulfur bacteria the special

pair of chlorophylls binding P870 has an oxidation potential of 0.45 V. The special pair is also

present in oxygenic photosynthesis in the core of the reaction center of PSI binding the two P700

and has an oxidation potential of 0.49 V. In contrast the reaction center of PSII contains a

tetramer of 4 chlorophylls known as P680, with an oxidation potential of 1.17V. Both PSI and

PSII systems are present in oxygenic photosynthesis in plants, algae and cyanobacteria (Barber

and Archer, 2001).

3.3.1 ANOXYGENIC PHOTOSYNTHESIS

Anoxygenic photosynthesis is carried out by a domain of bacteria groups which include, the

purple nonsulfur bacteria, the purple sulfur bacteria, the green sulfur bacteria and the

Heliobacteria. These bacteria use bacteriochlorophyll (a, b, c, d, e and g, depending of the

bacteria group) as the main light harvesting molecules and uses complimentary pigments, such as

carotenoids, to transfer the harvested energy to the bacteriochlorophyll. The bacteria have

chlorosomes where the light gathering system is located, the counterpart in plants are the

thylakoids located in the chloroplast where the light reaction and light independent reaction take

place (Staley et al., 2007).

3.3.1.1 THE PURPLE NON-SULFUR BACTERIA

The purple nonsulfur bacteria are found in lakes, sediments and ponds. They use hydrogen as an

electron donor (not water), and use the electrons from hydrogen to create a proton gradient

similar to the cyanobacteria in oxygenic photosynthesis. This proton gradient is used to drive

ATP synthase for production of ATP and these bacteria use a reverse electron transfer to produce

NADPH. The light independent reaction reduces CO2 to synthesize carbohydrates, however,

these bacteria are capable of using several pathways for carbohydrate production, one being the

Calvin cycle, however this pathway is not the main one as it is in oxygenic photosynthesis.

During anoxygenic photosynthesis water is synthesised as by-product (Staley et al., 2007).

Page 45: Effects of Light, Co2 and Reactor Design on Growth of Algae

31 

 

3.3.1.2 THE GREEN AND PURPLE SULFUR BACTERIA

The purple sulfur bacteria are similar to the purple nonsulfur bacteria: they use the same

bacteriochlorophyll a and b. However, the purple sulfur bacteria use hydrogen sulfide as an

electron donor and accumulate elemental sulfur in granules inside the cell, and is able to tolerate

high levels of sulfur that are otherwise toxic for the purple nonsulfur bacteria. In contrast, the

green sulfur bacteria possess a different type of bacteriochlorophyll c, d and e and another type

of carotenoids as complementary pigments. Pigments such as chlorobactene and isorenieratene

allow these bacteria to harvest more energy, and for this reason these type of bacteria do not

require the reverse electron transfer to produce NADPH which is characteristic of purple

bacteria. Another major difference is that the sulfur residues in the green sulfur bacteria are

synthesized outside the cell and not inside the cell like in purple bacteria (Staley et al., 2007).

3.3.1.3 HELIOBACTERIA

The heliobacteria are gram positive bacteria found in soils and contains a bacteriochlorophyll g

(788 nm) that has some similarities with the chlorophyll a present in plants. These bacteria do

not have chlorosomes and the light gathering system is attached to the cytoplasmic membrane.

These bacteria do not fix CO2 but use simple organic compounds (acetate, lactate and pyruvate)

thus perform photoheterotrophic growth, as carbon sources (Staley et al., 2007).

3.3.1.4 RHODOPSIN PHOTOSYNTHESIS

Rhodopsin is the type of photosynthesis present in the Archaea and bacteria domains growing

under environmental conditions where there can be both aerobic and anaerobic conditions. Their

photosynthetic machinery contains only bacteriorhodopsin (absorption peak 570 nm) with no

accessory pigments (they are the only photosynthetic organisms which do not have

complementary pigments). These organisms do not have an electron transport chain as do other

photosynthetic organisms, but have a light-powered proton pump which produces the proton

gradient necessary to drive ATP synthase (Staley et al., 2007).

Page 46: Effects of Light, Co2 and Reactor Design on Growth of Algae

32 

 

3.3.2 OXYGENIC PHOTOSYNTHESIS

The oxygenic photosynthesis mechanism provides most of the chemical energy essential for

living organisms. For eukaryotic organisms it includes plants, multicellular algae and

microscopic algae, and for bacteria includes the cyanobacteria group. The architecture of

photosynthesis in plants, algae and cyanobacteria is formed by two states of energy harvest:

photosystem I and II. Photosystems I and II share similar architecture in which both have a light

harvesting complex, known as antennas, containing chlorophyll and complementary pigments.

Each antenna has a reaction center (RC). In the antenna the photosynthetic pigments absorb

electromagnetic wavelengths in the visible spectrum and transfer this vibrational energy to the

RC. The vibrational energy charges an electron from a ground state to an excited state, and it

then moves through an electron transfer chain, releasing its free energy to produce ATP and

NADPH (Yan et al., 2003).

In oxygenic photosynthesis both eukaryotes and bacteria use chlorophyll as the main molecule to

harvest light and convert it to chemical energy. Both domains also use accessory pigments to

harvest light and to pass the energy to the chlorophyll a antenna also known as the chlorophyll a

RC. The complementary molecules are carotenoids, carotenes, xanthophylls, phycobillins and

their subdivisions (Campbell et al., 1999).

During oxygenic photosynthesis, H2O is oxidised (split into H+ + H+ + O-2, oxygen is released as

O2); the two electrons from water are energized twice, the first in the photosynthesis II complex

to produce the proton gradient that generates ATP; the second is in the photosynthesis I complex

to energyze the ferredoxin NADP+ reductase (FNR) producing NADPH from NADP+ + P. These

transformations are known as the light reaction and involve photosynthesis II (P680),

photosynthesis I (P700), cytochrome b6f, ferredoxin NADP reductase (FNR), ATP synthase and

the mobile plastoquinones Qa and Qb, plastocyanin, and ferredoxin (Campbell et al., 1999).

The photosystem I with P700 chlorophyll, discovered in 1957, has a signal formation of 10 ns

and flash decay time lasting 10-20 ms. PSII with P680 chlorophyll was discovered more recently

Page 47: Effects of Light, Co2 and Reactor Design on Growth of Algae

33 

 

as the flash decay lasted just 0.1-0.2 ms and was considered P700 noise in earlier studies. The

absorption peaks for P700 are 438 and 700 nm, whereas they are 435 and 680 nm for P680. The

oxidation potential of P680 is 1.17 V, which is greater than that necessary for the oxidation of

water (1.0 V). P700 has an oxidation potential of 0.49 V, chlorophyll a 0.78 V and carotenoid

1.06 V (Barber and Archer, 2001).

There is an additional reaction occurring during photosynthesis that is independent of the light

reaction and is known as the Calvin cycle. It uses 6 ATPs and 6 NADPHs to fix 3 CO2 molecules

and to produce 1 molecule of glyceraldehyde 3 phosphate (G3P). The Calvin cycle also requires

an extra 3 ATPs to regenerate the ribulose bisphosphate (RuBP) that is the CO2 acceptor. In total

9 ATPs and 6 NADPHs are used (Campbell et al., 1999).

The photosynthesis process has stabilized the fleeting charge separated state of energy into a

controlled chemical redox state thereby transforming electromagnetic radiation into chemical

energy. PSI in combination with PSII transfers the excited state electron from cytochrome C6 or

plastocyanin to ferredoxin and to ferredoxin NADP reductase to store the energy for generating

ATP by the PSII reaction whereas the energy required to sysnthesize NADPH from NADP is

from PSI. The free energy stored as ATP and NADPH are used to produce G3P during the light

independent reaction and this is the main product of photosynthesis (Di Donato et al., 2011;

Lubner et al., 2011). The PSI is located in the stroma lamellae and PSII is located in the grana

stacks. The light independent reaction takes place in the stroma of the chloroplast (Haldrup et al.,

2001).

3.3.3 ARCHITECTURE OF THE PHOTOSYNTHETIC APPARATUS IN PLANTS AND

ALGAE

The photosynthetic apparatus in eukaryotic cells such as plants and algae is located in a structure

called the chloroplast. Inside the chloroplast is an aqueous fluid, called stroma, and an

arrangemment of disks of 300-600 nm diameter size. These disks are called thylakoids and they

have an aqueous solution called lumen. The thylakoids are arranged in stacks of disks called

grana, and the network between grana is called stroma lamellae or lamella (Kouril et al., 2012).

See Figure 3.1 for details of chloroplast architecture.

Page 48: Effects of Light, Co2 and Reactor Design on Growth of Algae

34 

 

Figure 3.1. Chloroplast structure.

The photosynthetic apparatus is located in the thylakoid membrane, and the cytochrome 6bf is

located between the two membranes grana and stroma. PSI and PSII are located in the grana

where 80% of the photosynthetic pigments are present and carry out processes in the electron

transport chain. PSI and PSII are also present in the stroma lamellae where 20% of the total

photosynthetic pigments and activity are organized in a cyclic electron transport chain. PSI has

57-60% of the total photosynthetic pigments, therefore has the ability to capture more of the light

energy than does PSII (Albertsson, 2001). See Figure 3.2 for more detail of the thylakoid disk.

Page 49: Effects of Light, Co2 and Reactor Design on Growth of Algae

35 

 

Figure 3.2. Thylakoid disk including PSI, PSII and ATP synthase.

A grana stack is composed of 4 thylakoid disks and a stroma lamellae that links grana stacks.

Each thylakoid disk has a diameter close to 500 nm (appressed membrane and end membrane)

and width of 60 nm. Approximately there are 500 PSII and 250 cytochrome b6f located in the

appressed membrane, 210 PSI and 120 cytochrome b6f in the margin annulus, and in the stroma

lamellae there are 190 PSI, 50 PSII and 80 cytochrome b6f. See Figure 3.3 for a normal state of

transition of a thylakoid disk, however these ratios change depending of the pool sizes of ATP

and NADPH. (Albertsson, 2001).

During the photosynthetic process 6 hydrogen ions (6H+) are added to the lumen: two from the

oxidation of water, two from the activity of plastoquinones and two from cytochrome 6bf. The

activities of plastoquinones and the cytochromes are driven by the free energy obtained from the

two electrons charged during PSII. These two electrons are recharged in PSI to produce one

NADPH, a process organized in a linear electron transport chain. However, 3-4 or more

hydrogen ions are required to produce one ATP, and the light independent reaction requires 1.5

ATP for each NADPH produced. Therefore there is a cyclic electron transport chain between

PSII and PSI that pumps the remaining hydrogen needed to produce the ATP for the light

Page 50: Effects of Light, Co2 and Reactor Design on Growth of Algae

36 

 

independent reaction. The pool of ATP and NADPH or ATP:NADPH ratio induces changes in

the stroma lamellae including the PSI and PSII ratio, and the efficiency of PSI and PSII required

to harvest energy is also modified. These changes in amounts of PSI and PSII as well as the

efficiency of each structure to capture energy are known as state of transition (Albertsson, 2001).

Figure 3.3. Organization of the thylakoid disk including the stroma lamellae.

3.3.4 GENERAL DESCRIPTION AND FUNCTION OF PHOTOSYSTEM II (PSII)

Photosynthesis II (PSII) is a membrane-bound protein complex (700 KDa) located on the

thylakoid membrane, the protein complex is divided into 20 subunits which includes ~60 organic

and inorganic molecules. PSII is formed by two main structures, the antenna and a reaction

center (RC) (Service et al., 2011). PSII is formed from 27-28 proteins subunits, the antenna and

Page 51: Effects of Light, Co2 and Reactor Design on Growth of Algae

37 

 

the reaction centre (RC) are attached with CP47 and CP43 proteins. The RC contains 35-36

molecules of chlorophyll and 11-12 β-carotenes. The system is organized in 6 arrays, each one

with 5-6 chlorophylls and 1-2 β-carotenes (Alstrum-Acevedo et al., 2005).

The antenna is divided into a core antenna and a peripherical antenna; The core antenna

containing xanthophylls and light harvesting complex II (LHCII), the peripheral antenna

containing chlorophyll a and b, carotenoids, as well as 2-4 LHCII, subdivided into proteins;

Lhcb1 (66%), Lhcb2 (24%), Lhcb3 (8%) and minor complexes Lhcb4, Lhcb5 and Lhcb6 (Kouril

et al., 2012). The RC contains a core of two chlorophyll P680, chlorophyll D1 and chlorophyll

D2, D1 and D2 proteins, pheophytin D1 and D2, plastoquinones Qa and Qb, water oxidation

complex (WOC), which includes one redox active tyrosine (tyr161), the Mn4 cluster, oxygen

evolve complex (OEC) and the remainder of the electron transport chain (Alstrum-Acevedo et

al., 2005; Barber, 2002). The peripheral antenna functions in light absorbance, and in

transporting the vibrational energy to the core antenna. It also has a photo-protective mechanism

that dissipated excess energy as heat (Alstrum-Acevedo et al., 2005), the function of the reaction

center (RC) is to oxidize water (H2O), synthesis of oxygen (O2) and electron excitation

(energization) from the split of water (Kouril et al., 2012). There are two major routes for

transferring the vibrational energy from the antenna to the RC. One route is from the light

harvesting complex II (LHCII) attached protein CP26 to CP43 and to the RC. The other route is

from the light harvesting complex II (LHCII) attached protein CP24 to CP29, then to CP47 and

to the RC. Both routes from the antenna to the RC have to be through either CP43 or CP47

(Kouril et al., 2012). See Figure 3.4 for 3-D structure of the PSII antenna and RC.

Page 52: Effects of Light, Co2 and Reactor Design on Growth of Algae

38 

 

Figure 3.4. Three-dimensional structure of the PSII antennae and reaction center.

In cyanobacteria PSII has 35 chlorophyll a, 2 pheophytin and 12 carotenoids. Some of the PSII

carotenoids are xanthophyll, neoxanthin and lutein. There is a maintenance mechanism in PSII to

repair degradation of the D1 protein in the core of the RC. D1 proteins have a half life of ~30

min and about every 60 min have to be replaced with a new copy. PSII has two copies of the D1

protein so when a D1 protein does not function the other copy absorbs the excited-state electron

from the RC. Thus, the photosynthetic process does not stop during the repair cycle (Caffarri et

al., 2009).

The photosynthetic process in PSII initiates when energy is absorbed by the 35-36 chlorophyll

antennae and the 11-12 β-carotenes, and this excitation or vibrational energy is transferred to one

of the two chlorophylls P680 located in the reaction center. The vibrational energy excites an

electron inside of chlorophyll P680 and this excited state electron leaves the chlorophyll P680 or

chlorophyll RC through one of the chlorophyll D1/D2 proteins which reduces one of the

pheophytin D1/D2. The pheophytin passes the excited electron to the plastoquinone QA (mobile

Page 53: Effects of Light, Co2 and Reactor Design on Growth of Algae

39 

 

carrier), and QA reduces QB therefore transferring the energy of the excited electron.

Plastoquinone QB requires two excited electrons and two protons (2H+) from the stroma which

are released from PSII to a lipid bilayer known as a lumen. The process in PSII ends when QB

releases two protons into the lumen and transfers the two excited electron to the cytochrome b6f

complex (Barber, 2008; Ferreira et al., 2004).

3.3.4.1 PSII REACTION CENTER PROCESSES: P680, D1-D2 PROTEINS, PHEOPHYTIN,

OXYGEN EVOLUTION CENTER (OEC), WATER OXIDATION PROCESS (WOC) AND

PLASTOQUINONES

Photosystem II is a plastoquinone-oxidoreductase reaction that occurs inside the RC. It is driven

by light absorption in the antennae complex, which involves absorption of electromagnetic

energy, charging of an electron to an excited state, releasing this excited electron to cytochrome

6bf using the mobile carrier plastoquinones a and b, and resulting in water oxidation and oxygen

release. These processes can be separated into three reaction sequences:

1. Photon absorption by the antennae, transfer of this vibrational energy to chlorophyll P680

which excites an electron in the RC, and transfer of the excited electron to plastoquinone a and

plastoquinone b.

2. Oxidation of water which takes place in the water oxidation complex (WOC) using tyrosine

Yz, an intermediate redox carrier.

3. Reduction of plastoquinones which involves the transfer of an excited electron from the RC to

the plastoquinones using the heterodimer proteins D1 and D2. The electron flow is from the D1

and D2 proteins, which have a special pair of chlorophyll PD1 and PD2, chlorophyll D1 and D2

and pheophytin D1 and D2, to plastoquinone a and plastoquinone b (Renger and Renger, 2008).

The photosynthetic electron transport chain in the PSII RC occurs in the sequence: (i) charging

of the tetra dimer P680, (ii) excited electron flows to pheophytin (in a few picoseconds). (iii) The

electron flows from pheophytin to plastoquinones Qa (200 pico seconds) and one by one, two

Page 54: Effects of Light, Co2 and Reactor Design on Growth of Algae

40 

 

electrons flow to plastoquinone Qb (microseconds to milliseconds). And (iv) plastoquinone Qb is

converted to plastoquinol and transports the two electrons to cythochrome b6f. Water oxidation

and electron transport to P680 involves three steps: (i) water is oxidised by a Mn4 cluster which

has five internal redox states, S0 to S4, (ii) the redox- active tyrosine (tyr161) oxidizes the Mn4

cluster. And (iii) the P680 oxidizes tyr161 (10 nanoseconds) extracting the excited electron,

originally from water, and using it as an energy source to introduce the resonance energy

captured by the antenna (Barber and Archer, 2001).

Figure 3.5. Proposed structure of PSII based on X-ray images.

The PSII does not have 2 special pair of chlorophyll common in PSI, and anoxygenic

photosynthesis such as P840 and P870, therefore PSII cannot absorb light in the far red

spectrum. The general structure of the PSII RC (Figure 3.5) contains four chlorophylls- P680 and

2 pheophytin chlorophylls, with D1 and D2 proteins sharing the P680 chlorphyll and each

protein having one pheophytin. The D1 protein also has plastoquinone Qa, plastoquinone Qb, the

oxygen evolution center (OEC) and the water oxidation complex (WOC) which includes one

redox active tyrosine (tyr161) and the Mn4 cluster. In PSII the antenna is attached to the RC in

Page 55: Effects of Light, Co2 and Reactor Design on Growth of Algae

41 

 

the following sequence: D1 and D2 proteins are attached to CP47 protein, CP47 protein is

attached to cytochrome b559, cytochrome b559 is attached to PsbI, PsbTc, PsbW and then to the

antenna. P680 is located in the middle of the PSII RC and functions as its primary electron

donor. P680 in its oxidised state has a redox potential (1.12V) which more than sufficient to

oxidise water (Barber, 2002). PSII contains an oxygen evolving site, the Mn3-CaO4 cluster,

where the photooxidation of water occurs. The PSII contains 19 different proteins with 36

chlorophyll a in 3 subunits of 12 chlorophylls, 7 carotenoids, 1 reaction center (RC) with a pair

of chlorophylls, 1 oxygen evolving center (OEC), 1 heme b and 1 heme c, 1 plastoquinone a and

b, 2 pheophytins D1-CP43- PsbI and D2- CP47 -PsbX subunits. It has the following dimensions:

depth 105 A, length 205 A and width 110 A (Ferreira et al., 2004).

3.3.4.2 OXYGEN EVOLUTION CENTER (OEC)

The evolution of anoxygenic to oxygenic photosynthesis has changed the electron donor from

H2S or H2 to H2O and it involves the production of O2. The H2O oxidation reaction takes place in

the oxygen evolution complex (OEC) located inside PSII (Service et al., 2011). The PSII system

extracts electrons from water in the OEC and transfers these electrons to the P680 in the core of

the RC (Gilchrist et al., 1995). The OEC is located in the thylakoid membrane, nearby the PSII

RC. The WOC has a CaMn4 cluster (tetra nuclear clusters of manganese with calcium and

chlorine) where O-2 + O-2 couples and forms molecular oxygen O2 and it plays an active role in

proton distribution after the oxidation (Liu et al., 2008).

The OEC is activated by an excited electron transfer from the RC, received as vibrational energy

harvested by the chlorophyll antennae. The transfer of an excited electron from the RC to the

plastoquinones QA produces in the P680 sufficient oxidised equivalents to trigger water

oxidation in the OEC. This process uses the redox active tyrosine Yz as transport between the

chlorophyll and the OEC (Liu et al., 2008). The WOC and OEC, (they have similar structures)

are located in the D1 protein as a complex containing a tetra manganese cluster Mn4 with

chloride and calcium cofactors. It is in this structure that the oxidation of water takes place. The

transfer of electrons from the OEC to the P680 uses two tyrosines, a redox-active tyrosine Yz

(tyr 161) that transfers electrons from the manganese cluster to the P680 located in the D1

Page 56: Effects of Light, Co2 and Reactor Design on Growth of Algae

42 

 

protein. The other is tyrosine YD (tyr 160) which transfers electrons from the manganese cluster

to another P680 located in the D2 protein. The RC of PSII has 4 P680- two associated with the

D1 protein and two associated with the D2 protein. The location of the manganese cluster with

respect to the two tyrosine is not symmetric. The distance from the Mn4 to the Tyr 161 Yz is 4.5

Å both located in the D1 protein, and the distance from the manganese cluster to the Tyr 160 YD

is more than 4.5 Å (Gilchrist et al., 1995).

3.3.4.3 ARCHITECTURE AND DYNAMICS OF THE WATER OXIDATION PROCESS

(WOC)

The oxidation potential of P680 is 1.17 V which is greater than 1.0 V necessary for water

oxidation; P700 has an oxidation potential of 0.49 V, The oxidation potentical of chlorophyll a is

0.78 V, and the carotenoid is 1.06 V. In anoxygenic photosynthesis, P840 has an oxidation

potential of 0.4 V, P870 0.45 V and bacteriochlorophyll 0.64 V (Barber and Archer, 2001). In

photosystem II the water oxidation complex (WOC) architecture has 10 redox cofactors bound to

the D1/D2 proteins and the catalytic site consists of the CaMn4Ox cluster and the redox active

tyrosine Yz. The photosynthetic process is initiated in the antennae where electromagnetic

energy is collected and transferred to the P680, which is the primary electron donor located in

the RC and is made up of four chlorophylls. The P680 uses the energy transferred by the antenna

to charge an electron and donates this electron to D1 protein using chlorophyll D1 and

Pheophytin D1 as the electron transport chain. The P680 chlorophyll becomes considerable

oxidized after the loss of the excited state electron, this oxidation potential is strong enough to

extract a ground stage electron from the CaMn4Ox cluster in the core of the water oxidation

complex (WOC). The CaMn4Ox cluster involves four steps and five oxidation stages S0, S1, S2,

S3 and S4 where S0 is the most reduced state and S4 the most oxidised. P680 oxidise the

CaMn4Ox cluster and bring the oxidation from S0 to S4, when the WOC is at S4 state, in this stage

of oxidation the CaMn4Ox cluster has the potential to oxidise water extracting the electrons

locate in the hydrogen-oxygen bonds, water is split and the complex returns to S0 where the P680

start the oxidation process again oxidising the CaMn4Ox cluster. The electrons resulting from

water oxidation flow to P680 located in the RC. In the RC, vibrational energy from the antenna

Page 57: Effects of Light, Co2 and Reactor Design on Growth of Algae

43 

 

is used to excite an electron in the P680 and the electron transport chains starts over again in

PSII. The WOC provides a pH of 5 and a voltage difference of 0.93V per electron, for a total of

3.72 eV required to oxidise one water molecule (Chen et al., 2011; McEvoy et al., 2005).

Each state of oxidation in the Mn4Ca cluster is a combination of Mn3+ and Mn4+; the S0 state

spend 36% of the time and is the state after the oxidation of H2O and the release of O2; S1 state

spend 62% of the time and has 2 Mn3+ plus 2 Mn4+; S2 state spend 2% of the time and has 1

Mn3+ plus 3 Mn4+; S3 and S4 states are neglible most of the time and the formation of these

states triggers the oxidation of H2O and the release of O2 (Service et al., 2011).

3.3.4.4 MECHANISM OF ENERGY TRANSFER FROM PSII TO THE MOBILE

CARRIERS

The PSII complex absorbs oxygen, oxidizes water and reduces plastoquinones. PSII has two

branches located across a membrane that transfer the charged electron to plastoquinone, one

branch being active at any specific time. The active branch contains the pheophytin D1 (PheoD1)

protein which accepts two excited electrons and is the primary electron acceptor in the transfer

chain to plastoquinones QA and QB. In the other hand the Pheophytin D2 (PheoD2) coupled to a

central chlorine in the RC to assits with equilibrium of the excited RC state, thus Pheophytin D2

(PheoD2) is involved in the proper distribution of excitation energy throughout the chlorophyll

in the excite RC. PheoD1 is involve in the transferring of the excited electrons from the RC to

the Plastoquinones QA to QB that is the stable electron acceptor, and the PheoD2 is involved in

stabilizing the charges in the RC in the excite state (Perrine et al; 2011).

The RC in PSII for purple bacteria and oxygenic photosynthesis in plants, algae and

cyanobacteria contains 4 bacteriochlorophyll or chlorophylls and bacteriopheophytin or

pheophytin plus 2 B-carotenes between chlorophylls. These molecules are arranged in two

branches, with one branch active for charge separation and electron transfer at any one time. In

the purple bacteriun Rhodobacter sphaeroides, the light harvest 1 (LH 1) absorbs energy at a

Page 58: Effects of Light, Co2 and Reactor Design on Growth of Algae

44 

 

peak wavelength of 875 nm, the energy collected charges an electron in the RC which is then

transferred to bacteriochlorophyll on the active branch. There are two pathways in the

subsequent electron flow: (i) an active branch using bacteriopheophytin, and (ii) a special pair of

bacteriochlorophyll using the active branch. After an active branch is charged, it can not receive

additional energy from the antenna. The two pathways have different rates of electron transfer

and the one used depends of the intensity of light absorption and on the temperature of the

environment. This allows the RC to cope with stressful situations and be able to adjust the rate of

charged electron transfer to plastoquinone QA, a similar mechanism is present in the RC of

plants (Romero et al., 2010).

3.3.4.5 ROLE OF PLASTOQUINONES IN PSII (MOBILE CARRIERS)

Plastoquinone A (QA) is the primary electron receptor of PSII, it is responsible for the transfer of

the excited electron from the RC to plastoquinone B (QB) and then to cytochrome b6f. The

reduced state of plastoquinone A (QA) increases as the concentration of xanthophyll (a

complementary pigment) increases and the rates of epoxidation increase, and during this process

the light harvesting process decreases. At the same time the composition of the polypeptides

involved in the light harvesting process in the PSII RC are regulated by the redox state of

plastoquinone A (QA). Light intensity and temperature produce similar effects in controlling and

regulating xanthophyll production and the epoxidation state (Perrine and Sayre, 2011).

Two indicators have been identified responding to changes in light intensity and temperature

during the photosynthetic growth and adaptation of the algae Chlorella vulgaris. One is the

increase in the redox state of plastoquinones as light intensity increases, and the second

indication is the change in pH of the lumen in the thylakoid disk. There are two pathways

responsible for the transfer of excited electrons from the RC to plastoquinones QA, this being a

strategy to regulate the rate of excited electron transfer depending on light and temperature

conditions (Perrine and Sayre, 2011). An additional strategy for regulating photosynthetic rate is

through the PSII/PSI ratio (Melis, 1998). The PSII: PSI optimizes the flow of electrons from the

RC to cytochrome b6f and to plastocyanin, and then to the PSI complex (Wilson and Huner,

2000).

Page 59: Effects of Light, Co2 and Reactor Design on Growth of Algae

45 

 

3.3.4.6 DAMAGE AND REGENERATION OF PSII

The regeneration of the RC of PSII is processed every 30 to 60 min under normal conditions

caused by the oxidative damage from the splitting of water into O-2 + O-2 and the formation of O2

before it is released (Liu et al., 2008; Barber and Archer, 2001). This process occurs inside the

chloroplast of plants, algae and inside the PSII of cyanobacteria. The regeneration process is

focused on the D1 protein as it is this protein that is in contact with the oxygen produced in the

OEC. If the D1 protein becomes damaged, the P680 stops absorbing energy from the antenna

due to lack of electrons supplied from the WOC. The P680 also has an oxidation protection

mechanism, known as P680 tetramer carotenoid triplet quenching mechanism, in which B-

carotene quenches the P680 to protect the chlorophyll from oxidation by O2 (Barber and Archer,

2001).

The degradation process occurring during photo-damage develops first to the D1 protein, whose

synthesis is encoded by the specific gene, psbA. The synthesis of the D1 protein can be

accomplished only in the presence of light and there are additional proteins involved in the

process. Degradation of the D1 protein is common during the process of photosynthesis,

therefore the D1 protein is synthesized during a normal photosynthetic period (Mulo et al.,

2012).

There are two types of maintenance processes of the RC machinery: the first process is called the

novo assembly, that is the synthesis and replacement of functional sub-units damaged during

photosynthesis, the second process is PSII repair whereby the D1 protein is repaired. Enzymatic

degradation of the existing or damaged D1 protein occurs prior to the new copy of D1 replacing

the degraded one (Mulo et al., 2012).

Thermo-degradation of the D1 protein in the RC of PSII occurs when cells are subjected to

temperatures higher than their biological normal. During the normal photosynthetic process, D1

protein regeneration follows the following steps: (i) disassembly, enzymatic degradation and

removal of the damaged D1 protein, and (ii) synthesis and assembly of the new D1 protein at the

same location as the damaged D1 protein. Under normal conditions rates of degradation and

Page 60: Effects of Light, Co2 and Reactor Design on Growth of Algae

46 

 

regeneration are equal but this steady-state becomes uncoupled when there is a permanent

change in the normal temperature. An increase from the normal temperature increases the

electron transfer flux in plastoquinones Qa-Qb, and this increases the reactive oxygen species

(ROS) inside the photosynthetic machinery. A high concentration of ROS induces D1 protein

oxidation and degradation, thereby decreasing D1 protein concentrations, and this overall

reduces photosynthetic activity. An increase in temperature enhances photo-respiration which

causes additional damage to the photosynthetic machinery, especially to D1 regeneration

(Dinamarca et al., 2011).

An excess of electromagnetic radiation produces high amounts of ROS, the three principal

species being super oxide radical O2-, hydrogen peroxide H2O2 and hydroxyl radical OH-, and

this inhibits the synthesis of the novo protein necessary to repair the PSII machinery. Specifically

ROS reduces the rate of D1 protein synthesis thereby slowing the replenishing of the D1 protein.

ROS does not damage the photosynthetic machinery, however it decreases the regeneration and

repair rates of the machinery. ROS targets the PSbA genes -there are 3 PSbA genes- that encode

for the novo D1 protein synthesis. H2O2 targets the initiation and elongation in translation of

PSbA mRNAs, thus affecting the binding process of PSbA mRNAs to ribosomes. ROS not only

affects the synthesis of the D1 protein but also inhibits synthesis of the majority of the proteins in

the photosynthetic machinery in PSII. The most important ROS is formed in the PSI when there

is excess light. The first ROS formed is the super oxide radical O2-, part which forms H2O2 and

some forms the OH-. Hydrogen peroxide is the most abundant ROS and though it is the least

reactive it causes the most damage to the D1 protein regeneration process (Nishiyama et al.,

2001).

3.3.4.7 ROLE OF PSB27 IN THE REGENERATION OF PSII

Psb27 is one of the 20-23 proteins present in the PSII complex located in plants, algae and

cyanobacteria. The PS II performs the harvesting of light, the splitting of water, the excitation of

electrons from the energy obtained from light, and the reduction of plastoquinones. Psbp27 has a

molecular weight of 11000 daltons and is located in the thylakoid lumen. There is evidence that

plants able to synthetize Psbp27 recover faster from photo-inhibition damage than do those

Page 61: Effects of Light, Co2 and Reactor Design on Growth of Algae

47 

 

plants that cannot synthetize Psbp27. The PSII containing Psbp27 do not have Mn4Ca in the RC

water splitting region, but contain a D1 protein (D2 protein is absent). When there is a photo-

damage to the proteins of PSII in cyanobacteria, the bacteria synthetize Psbp27 to conduct

photosynthesis during the period of repair to damaged proteins. When the repair is completed

Mn4Ca and D2 protein is synthetized in the RC splitting water site and Psbp27 is inactivated.

The Psbp27 protein is used during the formation of the Mn4Ca complex and D2 protein in the

PSII RC and in the repairing process of the PSII complex caused by photo-damage. The presence

of Psbp27 is an evolutionary advantage for photosynthetic cells to be able to survive adverse

environmental conditions with excess of light and low temperature (Mabbitt et al., 2009).

3.3.5 GENERAL ARCHITECTURE AND FUNCTION OF PSI

The architecture of PSI consists of membrane proteins made up of 11-13 polypeptides and

subunits of these proteins, the most important of which are PSaA and PSaB. They bind the core

of the antenna with electron transfer cofactors of the RC, including the primary donor P700

pigments and the iron sulfur clusters FX. The antennae consist of three major pools containing

90-100 chlorophyll molecules and 22 carotenoids, and in plants are contained in a structure

called a light harvesting complex I (LHCI) (not found it in cyanobacteria). The LHCI is a

peripheral structure surrounding the complex containing the antennae and has 100 chlorophylls,

including 77 chlorophyll a and 23 chlorophyll b and 20 xanthophyll molecules. The antenna

including the light harvesting complex initiate the electron transfer process by first forming a

stable charge-separated state. The excited state electrons are then transferred via ferrodoxin to

the ferrodoxin NADP reductase (FNR) which synthetize NADPH (Di Donato et al., 2011; Yan et

al., 2003; Chitnis, 2001; Palson et al., 1998).

There are two models to explain the excited electron energy transfer dynamics in PSI: the first is

the trap-limited model and the second is transfer via a trap kinetic scheme. The stromal protein,

known as PSaC, binds the terminal iron clusters FA and FB, the latter having a relatively long half

life ~50 ms and a midpoint potential of -580 mV which is used to reduce protons to H2 (Lubner

et al., 2011).

Page 62: Effects of Light, Co2 and Reactor Design on Growth of Algae

48 

 

PSI has been described as a light-driven engine located in the thylakoid membrane of prokaryotic

cells like cyanobacteria or in the chloroplast of algae and plants. PSI is a plastocyanin-P700-

ferrodoxin redox system which functions to energize an electron and transfer it to ferredoxin

reductase (FNR) for the synthesis of NADPH. The PSI complex has 11-12 proteins in

cyanobacteria and 11-14 proteins in algae and plants, plus organic and inorganic cofactors.

The P700 RC is made up of a pair of chlorophylls a, which converts the vibrational energy by

exciting an electron. This excited electron leaves the P700 RC and flows to chlorophyll a (also

called A0) through to phylloquinones (A1) to the FX, then to the Fa and Fb cofactors. Fb is the

terminal electron donor and it reduces ferredoxin, a mobile carrier that provides the excited

electron for NADPH synthesis (Chitnis, 2001).

3.3.5.1 PSI REACTION CENTER CHLOROPHYLL P700

Chlorophyll P700 in the core of the PSI RC is made up of a pair of chlorophylls PA and PB and

located adjacent to an accessory pair of chlorophylls, referred to as a special pair, for a total of 4

chlorophylls. The RC has two functional branches A and B, each protected by a protein, and

refered to as branch PSaA and branch PSaB, respectively. Inside PSaA is chlorophyll PA and

attached to chlorophyll PA is a special pair a 1A. Special pair a 1A is attached to chlorophyll a

A0. Chlorophyll A0 is attached to phylloquinone A1 which in turn is attached to an iron sulfur

cluster ferredoxin FX. In total, the PSI complex is made up of 28 chlorophyll PA and 72

chlorophyll PB, 98 reside in the antenna and 2 are in the RC. The energy flow follows the

sequence: energy captured by the antenna is transmitted as vibrational energy to the RC, and this

energy excites an electron located in each chlorophyll P700. Chlorophyll P700 PA or PB donates

an excited-state electron to the special pair 1A or 1B, and the special pair transfers the electron to

chlorophyll A0 or B0, then its flows to phylloquinone A1 or B1. Phylloquinone A or B transfers

the electron to the iron sulfur cluster FX, and then on to ferredoxin reductase FNR where the

energy from the electron is used to synthesize NADPH (Saito and Ishikita, 2011). Figure 3.6

shows the structure of the photosystem I.

Page 63: Effects of Light, Co2 and Reactor Design on Growth of Algae

49 

 

Figure 3.6. Photosystem I structure.

Each of the 90-100 chlorophylls a of the antennae of PSI are separated by a distance ranging

from 7 to 16Å, and the distance is increase to 19Å from the central chlorophylls. The antennae

also have complementary pigments, 20 β-carotenes. The core of the antenna has 4 chlorophylls,

two chlorophyll cC and cC’ located at distance of 14Å from the central chlorophylls eC3 and

eC3’. The central chlorophyll transfers the vibrational energy to P700 located in the RC. The RC

has 6 chlorophylls, 2 phylloquinones A1 and A2 and three iron sulfur clusters. The core of the

RC contains a dimer of chlorophyll, known as P700, 2 special pair of chlorophyll A and B, and

two chlorophyll a’s. The six chlorophylls are known as the A0 primary electron donor. The

Page 64: Effects of Light, Co2 and Reactor Design on Growth of Algae

50 

 

excited electron flows from A0 to phylloquinones A1 and then on to the iron sulphur cluster

(4Fe-4S) in a sequence FX, FA and FB. They are located near to the stroma where ferredoxin

reductase (FNR) receives the excited electron and generates NADPH. From the time of energy

absorption by the antenna to the transfer of this energy to the RC takes 20-30 Ps (10-12 S) for

plants, cyanobacteria and algae. The time required to transfer vibrational energy between

chlorophyll in the antenna is 200 fs (10-15 S) (Palson et al., 1998). See Figure 6 for PSI

organization.

The ~98-100 chlorophylls of the antenna in PSI have an absorption peaks ranging from 660 to

700 nm, 8 of which have an absorption peak ~703 nm. The antennae are located 30-50 Å from

the RC. The RC has two active branches, PSaA and PSaB, which pass the excited state electrons

to ferredoxin FX, contrary to PSII where only one branch pass excited stated electron the D1

protein branch. The antenna excitation decay varies depending on the organism species, and in

the case of cyanobacteria lasts for 19-24 ps (Pico second 1*10-12 seconds). Under normal levels

of radiatation the time from light absorption in the antenna and transmission of the energy to

P700 takes 20-25 ps (open antenna 23.5 ps and closed antenna 24 ps), however under conditions

of high irradiance the time can be reduced to 4 ps. Excited electron transfer from P700 to

chlorophyll a A0 takes 1-3 ps and from A0 to phylloquinone A1 takes 20-50 ps 27 (Savikhin et

al., 2000).

3.3.5.2 REDOX POTENTIAL OF PSI

The redox potential of PSI has increased during biological evolution as evident by the redox

potential being different between P700 and P700+ among photosynthetic organisms. There is an

increasing trend in redox potential from simple prokaryotic organisms to more complex

eukaryotic organisms and multi cellular organisms. Prokaryotic cyanobacteria, the first oxygenic

organisms, have a P700 to P700+ redox potential between +398 and +454 mv. The PS1 redox

potential in red algae, a eukaryote, varies from +431 to +433 mv and in green algae is +469 mv.

Plants a multicellular organims have a redox potential of +470 mv which is the highest for the

PSI of any photosynthetic organism. All organisms with chloroplasts have a higher redox

potential in PSI than that of eukaryotic organisms without chloroplasts. In general, the redox

Page 65: Effects of Light, Co2 and Reactor Design on Growth of Algae

51 

 

potential in the RC of PSI P700 to P700+ is +398 mv to + 470 mv for the heterodimer; 2 P700,

the two special pair of chlorophylls, in the chlorophyll a A0 part of the RC has a redox of +800

mv (Nakamura et al., 2011).

3.3.6 MECHANISMS TO MINIMIZE PHOTO-DAMAGE IN PSI AND PSII

Intense light radiation can damage the pigments contained in the antennae of PSI and PSII. The

photosynthetic apparatus has mechanisms to dissipate excess energy as heat or to store it and use

it later for photosynthesis when less illumination energy is available. In plants the P700+ cation

radical acts as to cool the antenna by absorbing the longer wavelength radiation. In plants, 70-

90% of P700 can be in the P700+ state for managing the excess energy. In cyanobacteria the core

antenna contains chlorophylls, Chl 735 nm, Chl 711 nm, and Chl 714 nm, which absorb the

excess energy. The excess energy is transferred from chlorophyll to chlorophyll and then to the

P700 chlorophylls by resonance mechanisms (Shubin et al., 2008).

In a typical photosynthetic reaction, the rate of the process is controlled by the rate of electron

flow through the electron transport chain. The rate at which cytochrome b6f oxidises

plastoquinones regulates the flow of electrons that are transfered from PSII to PSI. If an antenna

absorbs more energy than the available electrons can carry from PSI to ferredoxin, the quenching

or heat dissipation mechanism is activated in the antenna or in the reaction center to avoid

damage to the machinery (Shubin et al., 2008).

3.3.7 STATE TRANSITION PSII AND PSI SYNTHESIS RATIO AND

PHOTOSYNTHETIC ACTIVITY

The products of the light reaction in photosystem I and II are NADPH and ATP. NADPH is

produced in the stroma of the chloroplast and its energy is derived from the flow of electrons

from PSI. ATP is synthetized by a proton gradient from the lumen to the stroma throught the

ATP synthase and the energy is derive from PSII. PSI and PSII have different architectures and

light absorption characteristics. Whereas both photosystems capture light in the blue and red

Page 66: Effects of Light, Co2 and Reactor Design on Growth of Algae

52 

 

regions, PSI absorbs more in the far red region with the result that it needs to transmit more

energy to the electron to produce NADPH. The light harvesting efficiency and quantity of PSI

and PSII complexes are constantly regulated to balance both the electron flow between the two

systems and the ratio of ATP and NADPH synthesis. Known as the state of transition, it is a

mechanism for balancing the PSI and PSII ratio occurring due to changes in the light intensity

and wavelength and optimizes the photosynthesis process. (Minigawa, 2011).

The state of transition process works with the phosphorylation of the LHCII in PSII or the

movement of LHCII from the PSII to PSI, and not additional synthesis of LHCII in PSII. This

process is a key component for the optimum performance of PSI and PSII. Depending on the

light spectra, the photosynthetic cell increases the efficiency of either of the two systems to

balance the electron flow for the production of ATP and NADPH. There are two states of

transition. During state 1 there is more fluorescence and more energy in the red region that is

absorbed by PSII, therefore there is more activity in PSII than PSI and the plastoquinone pool

spends longer in the reduced form. This activates the LHCII protein kinase by phosphorylation

moving LHCII from PSII to PSI and increasing the PSI light absorption efficency. During state 2

there is less fluorescence yield and more absorption in the far red region therefore the

plastoquinone pool is oxidised for longer periods the LHCII protein kinase is not as activated and

the LHCII synthesis remains in PSII and there is more activity in PSI than PSII. Thus this

mechanism is driven by the redox condition of the plastoquinone pool activating a protein that

moves LHCII from PSII to PSI (Haldrup et al., 2001).

Page 67: Effects of Light, Co2 and Reactor Design on Growth of Algae

53 

 

Figure 3.7. Photosynthetic apparatus in algae.

3.3.8 BACTERIA STATE OF TRANSITION

The antennae of purple non-sulfur bacteria Rhodobacter sphaeroides contain a light harvesting

complex 1 and 2 (LHC1, LHC2); the LHC1 has 24-30 bacteriochlorophylls that form an arc in

the RC. LHC1 also has a pair of plastoquinones Qa and a polypeptide Pufx. The function of Pufx

is to ensure the oxidation of plastoquinone Qa by stabilizing the oxidation state of plastoquinone

Qb. The oxidation of plastoquinone Qa is critical as plastoquinone Qb obtains the excited-state

electrons derived from the core of the RC through the bacterio-pheophytin-plastoquinone Qa and

transports the two electrons to cytochrome bc1. Without Pufx plastoquinone Qb would be

stabilized by LHC1 and this would prevent electron flow from plastoquinone Qa to

plastoquinone Qb and ultimately stop the photosynthesis process (Stahl et al., 2011).

Page 68: Effects of Light, Co2 and Reactor Design on Growth of Algae

54 

 

3.3.9 CONCLUSIONS

In oxygenic and anoxygenic photosynthetic ecept heiobacteria the organisms use a source of

hydrogen (H2, H2S or H2O) during the light reaction to transform electromagnetic energy into

chemical energy and store this energy as ATP and NADPH. The photosynthetic machinery

collects solar energy in its antennas and passes this energy to reaction centers where a ground

state electron passes to an excited state and leaves the RC. The energy from the excited-state

electron drives the entire photosynthetic process in the light reaction. During the light

independent reaction the ATP and NADPH formed in the light reaction are used to produce G3P.

Contrary to respiration, the end product of the photosynthetic process is the formation of G3P,

and not the consumption of these carbohydrates.

Page 69: Effects of Light, Co2 and Reactor Design on Growth of Algae

55 

 

3.4 REFERENCES

1. Albertsson, P. A. 2001. A quantitative model of the domain structure of the photosynthetic membrane.Trends Plant Scien. 6: 349-353.

2. Alstrum-Acevedo, J. H., Brennaman, M. K., Meyer, T. J. 2005. Chemical approaches to artificial photosynthesis II. Inorg. Chem. 44: 6802-6827.

3. Barber, J. 2008. Crystal structure of the oxygen-evolving complex of photosystem II. Inorg. Chem. 47:1700-1710.

4. Barber, J. 2002. P680: what is it and where is it. Biochem. 55: 135-138.

5. Barber, J., Archer, M. D. 2001. P680 the primary electron donor of photosystem II. J. Photochem. Photobiol. Chem. 142: 97-106.

6. Brock, T. 1979. Biology of microorganisms. 3th Edition. Prentice Hall. New Jersey, USA. 802 p.

7. Caffarri, S., Kouril, R., Kereiche, S., Boekema, E. J., Croce, R. 2009. Functional architecture of higher plant photosystem II supercomplexes. Embo J. 28: 3052-3062.

8. Campbell, N.A; Reece, J. B; Mitchell, L. G. 1999. Biology 5th. Addison Wesley Longman Inc. Don Mills, Ontario, Canada. 1175.

9. Chen, G., Allahverdiyeva, Y., Aro, E. M., Styring, S., Mamedov, F. 2011. Electron paramagnetic resonance study of the electron transfer reactions in photosystem II membrane preparations from Arabidopsis thaliana. Biochim. Biophys. Acta. 1807: 205-2011.

Page 70: Effects of Light, Co2 and Reactor Design on Growth of Algae

56 

 

10. Chitnis, P. R. 2001. Photosystem I function and physiology. Annu. Rev. Plant Physiol. Plant Mol. Biol. 52: 593-626.

11. Di Donato, M., Stahl, A. D., Stokkum, M. V., Grondelle, R. V., Groot, M. L. 2011. Cofactors Involved in light-driven charge separation in photosystem I identified by subpicosecond infrared spectroscopy. Biochem. 50: 480-490.

12. Dinamarca, J., Shly-kerner, O., Kaftan, D., Goldberg, E., Dulebo, A., Gidekel, M., Gutierrez, A., Scherz, A. 2011. Double mutation in photosystem II reaction centers and elevated CO2 grant thermotolerance to mesophilic cyanobacterium. Plos One. 6: 1-13.

13. Ferreira, K. N., Iverson, T. M., Maghlaoui, K., Barber, J., Iwata, S. 2004. Architecture of the photosynthetic oxygen-evolving center. Science. 303: 1831-1837.

14. Gilchrist, M.L., Ball, J. A., Randall, D. W., Britt, R. D. 1995. Proximity of the manganese cluster of photosystem II to the redox-active tyrosine Yz. Proceedings of the National Academy of Science. 92: 9545-9549.

15. Haldrup, A., Jensen, P. E., Lunde, C., Scheller, H. V. 2001. Balance of power- a view of the mechanism of photosynthetic state transition. Trends Plant Sci. 6: 301-305.

16. Kouril, R., Dekker, J. P., Boekema, E. J. 2012. Supramolecular organization of photosystem II in green plants. Biochim. Biophys. Acta. 1817: 2-12.

17. Kruse, O; Rupprecht, J; Mussgnug, J. H; Dismukes, C; Hankameer, B. 2005. Photosynthesis: a blueprint for solar energy capture for biohydrogen production technologies. Photochem. Photobiol. Sci. 4: 957-969.

18. Liu, F., Concepcion, J. J., Jurss, J. W., Cardolaccia, T., Templeton, J. L. 2008. Mechanism of water oxidation from the blue dimer to photosynthesis II. Inorg. Chem. 47: 1727-1752.

Page 71: Effects of Light, Co2 and Reactor Design on Growth of Algae

57 

 

19. Lubner, C. E., Heinnickel, M., Bryant, D. A., Golbeck, J. H. 2011. Wiring photosystem I for electron transfer to a tethered redox dye. Energy Environ. Sci. 4: 2428-2434.

20. Mabbitt, P. D., Rautureau, G., Day, K. L., Wilbanks, S. M., Eaton-Rye, J, J., Hinds, M, G. 2009. Solution structure of Psb27 from cyanobacterial photosystem II. Biochem. 48: 8771-8773.

21. McEvoy, J. P., Gascon, J. A., Batista, V. S., Brudvig, W. 2005. The mechanism of photosynthetic water splitting. Photochem. Photobiol. Sci. 4: 940-949.

22. Minigawa, J. 2011. State transition the molecular remodeling of photosynthesis supercomplex that controls energy flow in the chloroplast. Biochim. Biophys. Acta. 1807: 897-905.

23. Mulo, P., Sakurai, I., Aro, E. M. 2012. Strategies for psbA gene expression in cyanobacteria, green algae and higher plants from transcription to PSII repair. Biochim. Biophys. Acta. 1817: 247-257.

24. Nishiyama, Y., Yamamoto, H., Allakhverdiev, S. I., Inaba, M., Yokota, A., Murata, N. 2001. Oxidative stress inhibits the repair of photodamage to the photosynthetic machinery. Embo J. 20: 5587-5594.

25. Perrine, Z., Sayre, R. 2011. Modulating the redox potential of the stable electron acceptor, QB, in mutagenized photosystem II reaction centers. Biochem. 50: 1454-1464.

26. Renger, G., Renger, T. 2008. Photosystem II: the machinery of photosynthetic water splitting. Photosynth. Res. 98: 53-80.

27. Romero, E., Van Stokkum, I. H., Novoderezhkin, V. I., Dekker, J. P., Van Grondelle, R. 2010. Two different charge separation pathways in photosystem II. Biochem. 49: 4300-4307.

Page 72: Effects of Light, Co2 and Reactor Design on Growth of Algae

58 

 

28. Saito, K., Ishikita, H. 2011. Cationic state distribution over the P700 chlorophyll pair in photosystem I. Biophys. J. 101: 2018-2025.

29. Savikhin, S., Xu, W., Chitnis, P. R., Struve, W. S. 2000. Ultrafast primary processes in PSI from Synechocystis sp PCC 6803 roles of P700 and A0. Biophys. J. 79: 1573-1586.

30. Service, R. J., Yano, J., McConnell, I., Hwang, H. J., Niks, D., Hille, R., Wydrzynski, T., Burnap, R. L., Hillier, W., Debus, R. J. 2011. Participation of glutamate-354 of the cp43 polypeptide in the ligation of manganese and the binding of substrate water in photosystem II. Biochem. 50: 63-81.

31. Shubin, V. V., Terekhova, I, N., Kirillov, B. A., Karapetyan, N.V. 2008. Quantum yield of P700+ photodestruction in isolated photosystem I complexes of the cyanobacterium Arthrospira platensis. Photochem. Photobiol. Sci. 7: 956-962.

32. Stahl, A. D., Crouch, L. I., Jones, M. R., Stokkum, I. V., Grondelle, R. V., Groot, M. L. 2011. Role of Pufx in photochemical charge separation in the RC-LH1 complex from Rhodobacter sphaeroides - an ultrafast mid-ir pump probe investigation. J. Phys. Chem. 116: 434-444.

33. Staley, J. T; Gunsalus, R.P; Lory, S; Perry, J. J. 2007. Microbial life 2nd Edition. Sinauer Associates, Inc. Massachusetts, USA. 1066.

34. Wilson, K. E., Huner, N. P. 2000. The role of growth rate, redox-state of the plastoquinone pool and the trans-thylakoid pH in photoacclimation of Chlorella vulgaris to growth irradiance and temperature. Planta. 212: 93-102.

35. Yan, M., Damjanovic, A., Vaswani, H. M., Fleming, G. R. 2003. Energy Transfer in photosystem I of cyanobacteria Synechococcus elongates model study with structure-based semi-empirical Hamiltonian and experimental spectral density. Biophys. J. 85: 140-158.

Page 73: Effects of Light, Co2 and Reactor Design on Growth of Algae

59 

 

36. Mejia-Rendon, S. 2006. Design of an experimental bioreactor for measurement of microbial photosynthetic carbon dioxide fixation. Guelph, Ontario, Canada. University of Guelph. 253p.

37. Nakamura, A., Suzawa, T., Kato, Y., Watanabe, T. 2011. Species dependence of the redox potential of the primary electron donor p700 in photosystem I of oxygenic photosynthetic organisms revealed by spectroelectrochemistry. Plant Cell Physiol. 52-5: 815-823.

38. Zhu, X.G; Long S.P; Ort D.R. 2008. What is the maximum efficiency with which photosynthesis can convert solar energy into biomass. Curr. Opin. Biotechnol. 19: 153–159.

39. Zhu, X.G; Long, S.P; Ort, D.R. 2010. Improving photosynthetic efficiency for greater yield. Annu. Rev. Plant Biol. 61: 235–261.

Page 74: Effects of Light, Co2 and Reactor Design on Growth of Algae

60 

 

 

CHAPTER 4 GROWTH OF TWO MARINE ALGAE AND TWO FRESH

WATER ALGAE UNDER SOLAR AND LED ILLUMINATION

4.1 ABSTRACT

Growth experiments with two marine algae, Tetraselmis sp and Nannochloris sp, and two fresh

water algae, Chlorella vulgaris and Scenedesmus sp, were conducted to study the effects of five

light treatments on growth rate during a 20-24 d incubation. The light treatments were: daily

solar radiation, daily solar plus 12 h white light, 24 h blue LED light, 24 h red-blue LED light,

and daily solar plus 12 h red-blue LED light. Exponential growth was observed during the first

18 d of incubation and was followed by a stationary phase. Specific growth rates were highest in

treatments which included daily solar illumination, reaching a final population density of 107

cells mL-1 after 18 d and standing biomass of 1.200 mg L-1. ANOVA and T test at α = 0.05 in

cell density, cumulative growth and biomass production indicated a significant difference

between treatments that included daily solar radiation and those treatments that included 24 h

LED illumination. There was no significant difference in growth rate response to light treatment

between algal species.

Keywords: growth rate; algal biomass; population density; cell weight

4.2 INTRODUCTION

Cultivation of algae has gained importance because of the increased in demand for fossil fuels,

food and the increase of greenhouse gases resulted in global warming (Hill et al., 2006). Algae

have great potential as renewable source fuels, feedstock for fish and animals, food productions,

food supplements, antioxidants, and other important biochemicals (Pittman et al., 2011; Spolaore

et al., 2006).

Page 75: Effects of Light, Co2 and Reactor Design on Growth of Algae

61 

 

Algae could become an alternative for reduction of carbon dioxide emissions in power

generation plants, cement plants, and for atmospheric carbon dioxide sequestration (Fan et al.,

2007; Yun and Park, 1997). Algae growth rate is controlled by various factors including medium

nutrient content, temperature, pH, disturbance, illumination source and intensity. Algal species

have evolved growing under solar radiation, which provides energy in the ultraviolet (5%),

visible (43%) and infrared (52%) electromagnetic spectrum (Kruse et al., 2005). The visible light

spectrum is divided into 6 main colors: violet 380-450 nm, blue 450-495 nm, green 495-570nm,

yellow 570-590nm, orange 590-620 nm and red 620-750 nm. Not all wavelengths are equally

absorbed by the algae’s main photosynthetic pigments. Chlorophyll a and chlorophyll b

preferentially absorb blue and red light, while the other wavelengths are absorbed in lower

magnitude by the complementary pigment carotenoids (Matthijs et al., 1995).

Algae have been grown under both photoautotrophic and heterotrophic conditions in outdoor

ponds and in open and closed photo-reactors exposed to varying light wavelengths and intensity.

Photo-reactors have combined exposure to solar radiation during the day with exposure to white

light during the night (Ogbonna et al., 2001). Lately algae have been grown under LED

illumination which provides specific wavelengths that match the algae absorption peaks (Yeh

and Chung, 2009). Many experiments have studied algal growth exposed to specific wavelengths

of red, blue, yellow, green, and white light (Wang et al., 2007). Marine algae such as Tetraselmis

suecica, and Nannochloropsis sp have been study for the production of feed stock in aquaculture,

polyunsaturated fatty acids (PUFA) for food supplements (Abiusi et al., 2013), fatty acid

concentration and composition for the production of biodiesel (Bondioli et al., 2013; Park et al.,

2013). Different light spectra has been tested with T. suecica for cell biochemical composition,

cell size and biomass production. Cell size has been reduced, eicosapentaenoic acid and cell

motility increased while exposing the algae under red light for 9 d (Abiusi et al., 2013).

Studies in fresh water alga Scenedesmus sp and Scenedesmus obliquus have been focused on

maximization of biomass production, increased in lipid concentration and changes in lipid

composition for renewable fuel production (Nayak et al., 2013; El-Sheekh., 2013).

Page 76: Effects of Light, Co2 and Reactor Design on Growth of Algae

62 

 

Methodologies in biomass sonication to increase carbohydrates availability for hydrogen and

ethanol fermentation have been studied using the alga Scenedesmus sp (Jeon et al., 2013), The

studies have concluded that Scenedesmus sp has great potential for the production of renewable

fuels such as biodiesel, ethanol and hydrogen. A two stage growth methodology one for biomass

production and the second one with nitrogen deprivation for lipid accumulation has been tested

using the alga Chlorella protothecoides to increase its potential for biodiesel production (Li et

al., 2013), A novel methodology to process biomass from Chlorella vulgaris has been tested for

antioxidant extraction (luetin) for food supplements and for fatty acid synthesis necessary for

biodiesel production (Prommuak et al., 2013). This study examined the growth rates, cell

density, and biomass production of marine algae Tetraselmis sp, Nannochloris sp and fresh water

algae Scenedesmus sp, Chlorella vulgaris during a 20-24 d incubation under different

illumination treatments including solar radiation, white light, LED blue light and a combination

of red and blue LEDs. The objective of the research was to evaluate the effect of light on growth

for different algae species and to compare biomass production among marine and fresh water

algae growing under similar conditions.

4.3 MATERIALS AND METHODS

4.3.1 MICROORGANISM

The algae cultures used in this study were obtained from the algal culture collection located at

the University of Texas, Austin, TX. They included Chlorella vulgaris (26), Nannochloris sp

(1268), Scenedesmus sp (1589), Tetraselmis sp (2767).

Cultures were established in 1 L batch cultures for 15 d and then transferred to 4-L flasks. The

population density was measured using direct microscopic counting techniques. The population

density at the start of the experiment was measured to ensure accurate results of growth rate,

because the study was carried on with four different algae with different cell size, cell weight and

cell population. The reactors were set to 4 L at the beginning of the experiment.

Page 77: Effects of Light, Co2 and Reactor Design on Growth of Algae

63 

 

4.3.2 LIGHT SOURCE

4.3.2.1 EXPERIMENT 1 (DAILY SOLAR RADIATION)

For this experiment four 4 L reactors were set outside in an area exposed to only direct solar

radiation during the day and no artificial lighting during night period. The solar radiation

fluctuated from 8 µmol m-2s-1 at 6:00 pm to 2985 µmol m-2s-1 at 12:00 pm, the daily average

solar radiation measured during the course of the study was 1271 µmol m-2s-1. for all the

experiments light was measured using a Quantum meter (Model MQ, Apogee Instruments Inc,

Logan UT, USA).

4.3.2.2 EXPERIMENT 2 (DAILY SOLAR PLUS 12 H WHITE LIGHT)

For this experiment daily average solar radiation 1271 µmol m-2s-1 was supplemented after 6:30

pm until 6:00 am with white light provided by two 45 watt fluorescent cool white lamps

individual output 1850 µmol m-2s-1 located 10 cm from the reactors. In addition, one 65 watt cool

white lamp 2680 µmol m-2s-1 output was located 2 m above the 4 L reactors.

4.3.2.3 EXPERIMENT 3 (24 H BLUE LEDS)

This experiment was conducted in a room kept darkened. Blue illumination (425-435 nm) was

supplied with a 4.8 watt reflector containing 48 blue LEDs placed 15 cm from each reactor. The

illumination intensity measured at the reactor surface was 2850 µmol m-2s-1. This experiment

was conducted using 4 L reactors.

4.3.2.4 EXPERIMENT 4 (24 H RED + BLUE LEDS)

For this light treatment LED strips (LED 3m bands, ALAS, Shanghai, China) providing radiation

in the red 26 LEDs 0.08 W bulbs (620-625 nm) and blue 26 LEDs 0.1 W bulbs (425-430 nm)

wavelengths were used. The LEDs were contained in a clear plexiglass cylinder that was placed

Page 78: Effects of Light, Co2 and Reactor Design on Growth of Algae

64 

 

in the center of each 4 L reactor, in total 52 bulbs with a power consumption of 4.7 W were used

per reactor. Light intensity, measured using a Quantum meter, was 2143 µmol m-2 s-1 from the

blue LEDs and 1345 µmol m-2 s-1 for the red LEDs. The illumination system was supplied with

12 V and 8 amp power (Model SA 201-3485, ASTEC, California, USA).

4.3.2.5 EXPERIMENT 5 (DAILY SOLAR PLUS 12H RED AND BLUE LEDS)

Four 4 L reactors containing the four algae were placed in an outside area to receive direct solar

radiation averaging 1271 µmol m-2s-1 during the day which was supplemented with 12 h red and

blue LEDs during the night. LEDs were placed inside the reactor, the same type of LED

illumination was used as described in experiment 4.

4.3.3 LIGHT DISPERSION

The LEDs used in this study produced light emission intensity of 1700 to 2800 µmol m-2 s-1,

which is far greater than that required for optimal algal growth, which it is 50-250 µmol m-2 s-1,

there is a high probability of photo-inhibition, and even photo-damage, that will reduce the

photosynthetic activity of the cell (Kim et al., 2006; Degen et al., 2001; Ogbonna et al., 1999;

Merchuk et al., 1998; Fernandez-Sevilla et al., 1998; Molina –Grima et al., 1996; Garcia-

Sanchez et al., 1996; Qiang and Richmond, 1994; Aiba, 1982; Bannister, 1979; Eppley and

Coatsworth, 1996). Notwithstanding that the light emission from the LEDs is intense, light

dispersion occurs over short distances, and placement of the light source away from the reactor

reduces its intensity to a more favourable range for algal growth. The distance from the LEDs to

the algal suspension was 7 mm for the studies with red and blue light, and 10 cm for blue light

alone and 15 cm for the white light treatment receiving solar radiation. These distances were

sufficient to reduce the light intensity at the reactor surface to < 300 µmol m-2 s-1 (Ogbonna et

al., 1999).

Page 79: Effects of Light, Co2 and Reactor Design on Growth of Algae

65 

 

4.3.4 MIXING, QUANTIFYING CARBON SOURCE UNDER AUTOTROPHIC GROWTH

AND TEMPERATURE MEASUREMENTS

Five turbines (Model AC-9904, Resun, China) were used to inject bubbles with carbon dioxide

and to provide gentle mixing of the algal medium, 1125 mL air min-1 L-1 of algal suspension was

supplied with a CO2 concentration 0.0350% (350 ppm + 50) measured by a CO2-Temperature

monitor (Model 7001, Telaire-General Electric, California, USA).

Temperatures for outside experiments was 24°C + 3 for treatments 1, 2, 5 and 22°C + 0.9 for

treatment 3, 4 as measured in laboratory and outside conditions using a meter (Model 7001,

Telaire-General Electric, California, USA) and for liquid algal media as measured by an Infrared

thermometer (Model Fluke 62, Fluke Corporation, Everett, Washington, USA). Carbon dioxide

was measured every 30 min and temperature every h, outside treatments reached higher

temperature due to direct solar radiation into the reactors.

4.3.5 GROWTH MEDIA PREPARATION

The growth media for algal culture was prepared using a commercial synthetic fertilizer (Solucat

25-5-5, Atlantica Agricola, Villena, Spain), which contains total nitrogen - 25% (as ammonium-

14,2% plus nitrates-10.8%), P2O5 - 5%, K2O - 5%, Fe - 0.020%, Mn - 0.010%, Zn - 0.002%, B -

0.010%, and Cu - 0.002%. The medium for both the fresh water and marine algae contained 0.8

g fertilizer/L, initial electrical conductivity of the medium was 0.95 ± 0.01 mS measured by EC

monitor (Model HI 991301, Hanna, Michigan, USA). 35 g/L of sodium chloride salt was added

to the medium culture for the two marine algal species to simulate marine conditions at 3.5%

salinity. Volume culture were corrected by adding sterilized fresh media every time a sample

was taking, initial volume and final volume of the experiments were the same due to volume

correction.

Page 80: Effects of Light, Co2 and Reactor Design on Growth of Algae

66 

 

4.3.6 CELL DENSITY ANALYSIS

The alga cell density was measured daily throughout the time of the experiments by direct

microscopy using the Neubauer chamber. The methodology used is describe in 8 steps: 1) Shake

each algae reactor to make sure homogenous algal distribution before taking a 20 mL sample

with syringe, 2) for very concentrate sample dilute the sample with distillate water with a 1 to 10

ratio, 3) fill the Neubauer chamber with algal media, 4) depending of the cell size use the

microscopy 10X or 40X lens, 5) count the cells in the chambers A and B, for the small cells

(smaller than 6µm) count those cells in the chamber C, 6) for cells in chamber A and B use this

equation C = N * 104 * dil, where C = cells mL-1, N = cells average in 1mm2 (1µL) and dil =

dilution factor, 7) For small cells use the results in chamber C and this equation; C = (N/4*10-

6)*dil, where C = cells mL-1, N = cell average in the 5 C chambers and dil = dilution factor, 8)

Repeat steps 1 to 7 three times in order to have three repetitions.

4.3.7 GROWTH RATE METHODOLOGY

The growth of the algae during the incubations was calculated using the specific growth rate

equation K = ln (N2/N1)/(t2-t1), where K is the specific growth rate constant (h-1), N1 is the

algal population at time 1 (cells mL-1), N2 is the algal population at time 2 (cells mL-1), T1 is the

time at the start of exponential growth (h), T2 is the time after a period of exponential growth

(h). Divisions per day D = (K’/Ln2), Generation Time (Doubling time) GT = 1/D. Accumulative

or cumulative growth CG = (Total K/Ln2) (Andersen, 2005; Levasseur et al 1993; Dykhuizen,

1983).

4.3.8 BIOMASS AND CELL WEIGHT METHODOLOGY

The dry weight methodology was used for biomass quantification, this methodology is divided in

7 steps: 1) Dry a 15 mL glass tubing until reach constant weight, 2) fill out the tubing with 10

mL of the alga media from the experiment, 3) centrifuge the tubing at 4000 rpm for 15 min, 4)

pour off the water in the vial, 5) dry the tubing (105 oC) with the algae in an oven until reaches

constant weight, 6) weigh the tubing with the algae and subtract the weight of the tubing without

Page 81: Effects of Light, Co2 and Reactor Design on Growth of Algae

67 

 

the algae, and 7) the resulting value is the alga biomass in a volume of 10 mL. This procedure

was repeated in duplicate each time alga biomass was measured. Cell weight was calculated for

each alga by dividing biomass (g L-1) by cell population (cells L-1).

 

4.3.9 EXPERIMENTAL DESIGN AND STATISTICAL ANALYSIS

The study was set up using 20 bubble-column reactors (12 cm internal diameter and 40 cm in

length), each containing 4 L of growth media and exposed to five light treatments. Four different

algal species were used for each light treatment. Triplicate samples for cell density were

collected daily, measurements of final standing biomass were taken at the end of the experiment

in triplicate. Cell weight was calculated from the measurements of cell density and standing

biomass. ANOVA was performed using Biostatistics 1.0 at α≤ 0.05 for cumulative growth rate,

final population density of five different light treatments and four algae species, and for biomass

production of four algae growing under daily solar + 12h red-blue LEDs and 24h red-blue LEDs.

T test was used to compare differences among the specific light treatments.

 

4.4 RESULTS AND DISCUSSION

Figures 4.1 to 4.5 show the growth curves for the four different algal species growing under five

different light treatments. Most curves showed a sigmoid growth pattern and completed their

exponential growth stage within the 15 d incubation period and started the stationary stages after

approximately 16 d. Exceptions were those algae growing under daily solar radiation or growing

under 24 h LED red and blue. The highest population densities were achieved with those

treatments that combined solar radiation with exposure to a artificial light treatment. Growth

pattern of the two marine algae and two fresh water algae are similar for any specific light

treatment.

 

Page 82: Effects of Light, Co2 and Reactor Design on Growth of Algae

68 

 

4.4.1 ALGAL GROWTH UNDER DAILY SOLAR RADIATION

Figure 4.1. Algal growth under daily solar radiation.  

4.4.2 ALGAL GROWTH UNDER DAILY SOLAR RADIATION + 12 H WHITE LIGHT

Figure 4.2. Algal growth under daily solar radiation + 12 h white light.

Page 83: Effects of Light, Co2 and Reactor Design on Growth of Algae

69 

 

4.4.3   ALGAL GROWTH UNDER 24 H BLUE LEDS IN 4 L REACTORS 

Figure 4.3. Algal growth under to 24 h blue LEDs in 4-L reactors.  

4.4.4 ALGAL GROWTH UNDER 24 H BLUE + RED LEDS

 

Page 84: Effects of Light, Co2 and Reactor Design on Growth of Algae

70 

 

Figure 4.4. Algal growth under 24 h red + blue LEDs.

4.4.5 ALGAL GROWTH UNDER DAILY SOLAR RADIATION + 12 H BLUE +RED LEDS

Figure 4.5. Algal growth under daily solar radiation + 12 h blue +red LEDs.

4.4.6 FINAL POPULATION DENSITY FOR FOUR ALGAL SPECIES GROWING UNDER 5

LIGHT TREATMENT

The population densities developed by the marine algae during the course of this experiment was

6*106 to 9*107 and are in agreement with those previously reported with other algae species.

Studies with the marine alga Isochrysis galbana reached a population density of 3*107 cells mL-1

after 12 d of incubation (Fidalgo et al., 1998). In experiments with the marine alga Chaetoceros

calcitrans in different types of closed photo-bioreactors, a population density of 9*107 cells mL-1

was measured (Krichnavaruk et al., 2007). Experiments with Nannochloropsis sp reported

population densities of 1*108 Cells mL-1 (Fabregas et al., 2004) and with Tetraselmis suecica

densities of 1.3*106 and 7.8*106 cells mL-1 (Fabregas et al., 1985; Fabregas et al., 1984). In

Page 85: Effects of Light, Co2 and Reactor Design on Growth of Algae

71 

 

commercial production the marine alga Dunaliella salina has attained a population density in the

order of 109 cells mL-1 (Borowitzka, 1999).

Studies of fresh water algae have reported population densities in the same magnitude to those

presented in this study 1.6*106 to 3*107. Experiments with the red fresh water alga Porphyridium

purpureum has reached population density of 3*107 cells mL-1 in 16 d of incubation (Baquerisse

et al., 1999). Studies with the fresh water alga Chlorella ellipsoidea has reached population

density of 1.3*108 cells mL-1 (Cho et al., 2007). Experiments with the fresh water algae

Haematococcus pluvialis have reported a population density of 3.9*106 cells mL-1 (Garcia-Malea

et al., 2006).

Table 4.1. Population densities (cells mL-1) after 18 days of incubation for four different algae under five different light treatments.

Lighttreatmnet/Algae DailysolarDailysolarand12hwhitelight 24hBlueLEDs

Dailysolarand12hLEDs(RedandBlue)

24hLEDs(RedandBlue)

Tetraselmissp 2.65E+07 5.36E+07 8.96E+06 6.55E+07 2.49E+07Nannochlorissp 4.74E+07 7.14E+07 9.17E+06 5.69E+07 2.13E+07Scenedesmussp 8.39E+06 2.02E+07 5.46E+06 1.39E+07 8.65E+06Chlorellasp 1.06E+07 1.85E+07 1.60E+06 3.17E+07 7.80E+06

Mean 2.32E+07 4.09E+07 6.30E+06 4.20E+07 1.56E+07Standardeviation 1.80E+07 2.59E+07 3.56E+06 2.36E+07 8.69E+06

CoefficientofVariation 0.775 0.634 0.566 0.563 0.556

Finalpopulationdensity(CellsmL‐1)offouralgaeunderfivelighttreatments

Independent of the algal species, light treatments that included daily solar radiation reached

population densities in the magnitudes of 107 cells mL-1, see Table 4.1. The five light treatments

were tested for ANOVA and no significant differences were found; F = 2.9826 and p = 0.0537.

However the ANOVA test that include 4 treatments; all 24 h experimnets including daily solar

plus artificial ilumination and 24 h artificial illumination resulted in statistical differences with F

= 3.9358 and p = 0.0362. Individual treatmens were tested with the T tets analysis of variance;

Statistical differences were found in 24 h blue LEDs versus daily solar + 12 h cool white light

with T = 2.643 and p = 0.038. A significant difference was found in 24 h blue LEDs versus daily

Page 86: Effects of Light, Co2 and Reactor Design on Growth of Algae

72 

 

solar + 12 h red and blue LED with T = 2.993 and p = 0.024. The T test 24 h blue LED against

daily solar radiation showed no significant difference with T = 1.842 and p = 0.115. Further, the

T test showed that there is no statistical difference in daily solar radiation versus any treatments

that included daily solar illumination.

The effects of the different light sources on algal growth can also be examined with reference to

their specific growth rates as calculated during the exponential growth phase. Although the

growth rate and the cumulative growth rate is an important measurement, factors such as the

initial population, natural cell size of alga, turbulence of the growth media can affect the specific

growth rate. Another measure of the effectiveness of the treatments is standing biomass of the

final population density. For the four algae, Tetraselmis sp, Nannochloris sp, Scenedesmus sp

and C. vulgaris the best performance in terms of population densities were achieved when grown

under a combination of solar radiation (12 h) + artificial light for an additional 12 h. Under these

conditions, all four algae reached final population densities in the order of 107 cells mL-1 after 18

d growth. From the results of this study solar radiation plays a key role on the growth of high

population density in algae, mainly because this illumination contains all photosynthetic spectra

390-700 nm plus near infrared and provides the alga with sufficient energy reaching 3000 µmol

m-2s-1 with a daily average of 1271 µmol m-2s-1. However solar radiation is only available half of

the time, therefore when complemented resulted in a higher population density, as seen in the

results of this study. The two light treatments tested as complements of solar radiation increased

population density, however there was no conclusive difference which light complement

performed best.

4.4.7 SPECIFIC GROWTH RATE FOR VARIOUS LIGHT TREATMENTS

The marine algae, Tetraselmis sp and Nannochloris sp, obtained higher final populations under

the light treatments that included daily solar radiation, reaching a population density in the order

of 107 cells mL-1, and were similar in the white light and red + blue LED treatments. Tetraselmis

sp developed the highest cumulative growth rate of 7.6 in 18 d under treatments consisting of

daily solar radiation and 12 h white light. Though the initial population for this treatment was the

Page 87: Effects of Light, Co2 and Reactor Design on Growth of Algae

73 

 

lowest of all the five treatments, Nannochloris sp had a similar result under the same light

treatment, the cumulative growth rate was 7.4 in 18 d and the initial population for this treatment

was not the lowest. Tetraselmis sp attained the highest population density under exposure to 12 h

solar radiation plus 12 h red (625 nm) + blue (430 nm) LEDs. The final population density of

this treatment was 6.55 *107 cells mL-1. Under exposure to daily solar radiation plus 12 h white

light, Nannochloris sp reached a final population density of 7.1*107 cells mL-1, which was the

highest of the four algae. See Table 4.2 for more details

 

Table 4.2. Growth rate (d-1) under different lights treatments after 18 d of incubation.

Parameter Daily SolarDaily solar and 

12h white light24h Blue LEDs

Daily solar and 12h 

LEDs (Red and Blue)

24h LEDs         

(Red and Blue)

Initial population cells mL‐1 9.00E+05 2.75E+05 7.86E+05 6.60E+06 3.60E+06

Final population cells mL‐1 2.65E+07 5.36E+07 8.96E+06 6.55E+07 2.49E+07

Mean 0.220 0.371 0.170 0.146 0.133

Max 0.656 1.306 0.693 0.540 0.578

Min 0.050 0.004 0.000 0.001 0.020

Total in 18 days 3.364 5.272 2.886 2.295 1.932

Cumulative growth 4.854 7.605 4.163 3.311 2.787

Growth rate (days) of Tetraselmis sp  under five light treatments 

Parameter Daily SolarDaily solar and 

12h white light24h Blue LEDs

Daily solar and 12h 

LEDs (Red and Blue)

24h LEDs         

(Red and Blue)

Initial population cells mL‐1 2.88E+05 4.25E+05 3.25E+06 6.65E+06 4.75E+06

Final population cells mL‐1 4.74E+07 7.14E+07 9.17E+06 5.69E+07 2.13E+07

Mean 0.345 0.363 0.062 0.133 0.089

Max 1.739 0.809 0.175 0.352 0.290

Min 0.048 0.004 ‐0.009 0.016 0.000

Total in 18 days 5.083 5.124 1.037 2.147 1.498

Cumulative growth 7.333 7.392 1.496 3.097 2.161

Growth rate (days) of Nannochloris sp  under five light treatments

Page 88: Effects of Light, Co2 and Reactor Design on Growth of Algae

74 

 

Parameter Daily SolarDaily solar and 

12h white light24h Blue LEDs

Daily solar and 12h 

LEDs (Red and Blue)

24h LEDs         

(Red and Blue)

Initial population cells mL‐1 1.41E+06 3.50E+05 9.83E+05 1.85E+06 2.00E+06

Final population cells mL‐1 8.39E+06 2.02E+07 5.46E+06 1.39E+07 8.65E+06

Mean 0.104 0.254 0.096 0.121 0.098

Max 0.597 0.991 0.223 0.318 0.377

Min ‐0.815 0.011 0.019 ‐0.018 ‐0.006

Total in 18 days 1.772 4.053 1.714 2.013 1.464

Cumulative growth 2.556 5.847 2.473 2.904 2.113

Growth rate (days) of Scenedesmus sp  under five light treatments

Parameter Daily SolarDaily solar and 

12h white light24h Blue LEDs

Daily solar and 12h 

LEDs (Red and Blue)

24h LEDs         

(Red and Blue)

Initial population cells mL‐1 2.08E+06 5.25E+05 6.25E+05 4.55E+06 1.70E+06

Final population cells mL‐1 1.06E+07 1.85E+07 1.60E+06 3.17E+07 7.80E+06

Mean 0.101 0.210 0.033 0.123 0.108

Max 1.063 1.035 0.396 0.410 0.405

Min ‐1.474 0.031 ‐0.105 ‐0.002 0.010

Total in 18 days 1.635 3.549 0.940 1.940 1.523

Cumulative growth 2.359 5.121 1.356 2.798 2.198

Growth rate (days) of Chlorella sp  under five light treatments

The two fresh water algae, Scenedesmus sp and C. vulgaris, followed similar growth patterns.

For Scendesmus sp, the treatments that included solar radiation plus artificial lighting reached

population densities of 107 cells mL-1. Other treatments reached population densities of 106 cells

mL-1. The highest growth rates were recorded for algae exposed to daily solar radiation + 12 h

white light, reaching a final population density of 2.2*107 cells mL-1 and a cumulative growth

rate of 5.8 in 18 d. However, the initial population of this treatment was the lowest of all five

treatments. For C. vulgaris the best treatments was daily solar radiation + 12 h red and blue

LEDs. The final population of this treatment was 3.17*107 cells mL-1 and the cumulative growth

rate 2.8 in 18 d. The highest cumulative growth rate, 5.1, was obtained under daily solar

radiation + 12 h white light, however this treatment, as well as the treatment with Scenedesmus

sp, had the lowest initial population density of all treatments.

Page 89: Effects of Light, Co2 and Reactor Design on Growth of Algae

75 

 

The ANOVA test of the cumulative growth rate for five light treatments and four algal species

resulted in significant differences with F = 7.0543 and p = 0.0021. The individual T tests for

specific treatments showed significant differences with 99% confidence in 4 of 10 possible tests:

The light treatment daily solar + 12 h white light against 24 h blue LEDs resulted in significant

differences T = 4.669 with p = 0.003. Another significant difference was for daily solar + 12 h

white light against 24 h red and blue T = 6.707 with p = 0.000. Daily solar + 12 h white light

against daily solar + 12 h red and blue LEDs showed a significant differences T = 5.656 with p =

0.001. Significant difference were found for the light treatment daily solar + 12 h red and blue

LEDs against 24 h red and blue LEDs with T = 3.647 and p = 0.011. Overall the light treatments

affected significantly the growth and cell population of alga, the best treatment in cell density

and growth rate for two marine and two fresh water algae was the combination of daily solar +

12 h white light.

4.4.8 BIOMASS PRODUCTION

The standing biomass of alga growing under 24 h red-blue LEDs ranged from 910 mg L-1 to

1160 mg L-1, whereas alga growing under daily solar + 12 h red-blue LEDs developed biomass

in the range of 1090 mg L-1 to 1250 mg L-1. See Figure 4.6 for details.

Page 90: Effects of Light, Co2 and Reactor Design on Growth of Algae

76 

 

Figure. 4.6. Standing biomass of four alga species growing under two light sources.

Algal species growing under daily solar + 12 h red-blue LED illumination developed a higher

standing biomass than those alga species growing under 24 h red-blue LEDs alone. There was a

statistically significant difference in standing biomass F = 7.0049, with P = 0.0382 (more than

96% of confidences) between these two light treatments in the four alga species.

4.4.9 CELL WEIGHT

Algal cells weight ranged from 18 to 130 pico grams (pico gram 1*10-12 g), and differed between

algal species. The results show that algal cells are heavier when grown under 24 h LEDs

illumination than under daily solar + 12 h LEDs. The algae Tetraselmis sp and Nannochloris sp

had double the weight growing under the light source of 24 h LEDs; for C. vulgaris the

difference in weight was 3 times higher when the alga were grown under 24 h LED illumination

in comparison to daily solar + 12 h LEDs. See Table 4.3 for details.

Page 91: Effects of Light, Co2 and Reactor Design on Growth of Algae

77 

 

Table 4.3. Cell weight of four algae species growing under two light sources.

Cell weight (g)

Algae/Light source 24 h red-blue

LEDs Daily solar + 12 h red-blue LEDs

Tetraselmis sp 3.90E-11 1.77E-11 Nannochloris sp 5.46E-11 2.20E-11 Scenedesmus sp 1.05E-10 7.87E-11

Chlorella sp 1.29E-10 3.92E-11

The results of this study show that alga cell weight is altered by light treatments. A study

conducted with the alga Tetraselmis suecica under different LEDs showed that cell weight was

different when alga were grown under blue + red illumination (Abiusi et al., 2013). An

explanation for greater alga cell weight under 24 h LEDs and higher cell population and biomass

under daily solar + 12 h LED may be because of availability of space. At high cell populations,

culture space for growth in volume becomes limiting, therefore cell weight is smaller. Space

efficiency is better used by packing more cells per volume resulting in less free space and

therefore, more biomass. Another explanation may be to avoid mechanical damage as in a highly

populated system there is more possibility for cell collisions than in a less populated culture. In

addition smaller cell size would sustain less mechanical damage during collisions than would

bigger cells. From the observation of this study, the standing biomass and cell density is high

when alga cell weight of is low.

4.5 CONCLUSIONS

The nature of the light source used for illumination in algal reactors has a significant effect on

algal growth rate, cell density, biomass production and cell weight. The results of this study

confirm that light treatments affect these parameters in both marine and fresh water algae

species.

Page 92: Effects of Light, Co2 and Reactor Design on Growth of Algae

78 

 

We hypothesized that growth rate would be higher in those treatments exposed to 24 h

illumination compared to daily solar radiation. However, this study showed that a critical

component of the light treatment is exposure to solar radiation. Light treatments that included

solar radiation performed better than those that didn’t. The light treatments of daily solar

radiation performed better than 24 h blue LED and 24 h red + blue LED, presumably because

solar radiation provides all the light wavelengths required for algae growth.

Independent of algal species, the light treatments that produced the highest cell density, growth

rate and standing biomass are those that combined daily solar radiation with 12 h artificial

illumination during the night period. In these treatments degradation of algal biomass during the

night period due to algae respiration (Ogbonna et al., 1999; Ogbonna and Tanaka, 1996) are

minimized because photosynthetic processes are kept active during night time.

Alga cell weight is higher and standing biomass is lower in those algae growing under 24 h

LEDs illumination. As a result of the low cell population there is more space for alga cells to

grow, and this free space also reduced the probabilities of cell collisions of large heavy cells and

causing less mechanical damage, however space efficiency is low because there is a lot empy

space that can be filled with algae.

Further studies which combine 12 h LED illumination with daily solar radiation should be

conducted to establish a clear dependence of solar radiation as a key factor to ensure more

efficient algal growth. This study showed that daily solar radiation performed better than 24 h

exposure to artificial light.

Page 93: Effects of Light, Co2 and Reactor Design on Growth of Algae

79 

 

 

4.6 REFERENCES

1. Abusi, F., Sampietro, G., Marturano, G., Biondi, N., Rodolfi, L., De-Ottavio, M., Tredici, M.R. 2013. Growth photosynthetic efficiency and biochemical composition of Tetraselmis suecica F&M-M33 grown with LEds of different colors. Biotechnol. Bioeng. 9999: xx-xx.

2. Aiba, S. 1982. Growth kinetics of photosynthetic microorganisms. Adv. Biochem. Eng. J. 23: 85-156.

3. Andersen, R.A. 2005. Algal Culturing Techniques, p. 1-578. Elsevier Academic Press. China.

4. Bannister, T.T. 1979. Quantitative description of steady state, nutrient-saturated algal growth, including adaptation. Limnol. Oceanogr. 24: 76-96.

5. Baquerisse, D., Nouals, S., Isambert, A., Ferreira dos Santos, P., Durand, G. 1999. Modelling of a continuous pilot photobioreactors for microalgae production. J. Biotechnol. 70: 335-342.

6. Bondioli, P., Bella, L.D., Rivolta, G., Zittelli, G.C., Bassi, N., Rodolfi, L., Casini, D., Prussi, M., Chiaramonti, D., Tredici, M.R. 2012. Oil production by the marine microalgae Nannochloropsis sp F&M-M24 and Tetraselmis suecica F&M-M33. Bioresour. Technol. 114: 567-572.

7. Borowitzka, M.A. 1999. Commercial production of microalgae: ponds, tanks, tubes and fermenters. J. Biotechnol. 70: 313-321.

8. Cho, S.H., Ji, S.C., Hur, S.B., Bae, J., Park, I.S., Song, Y.C. 2007. Optimum temperature and salinity conditions for growth of green algae Chlorella ellipsoidea and Nannochloris oculata. Fish. Scien. 73: 1050-1056.

Page 94: Effects of Light, Co2 and Reactor Design on Growth of Algae

80 

 

9. Degen, J., Uebele, A., Retze, A., Schmid-Staiger, U., Trosch, W. 2001. A novel airlift photobioreactor with baffles for improved light utilization through the flashing light effect. J. Biotechnol. 92: 89-94.

10. Dykhuizen, D.E., Hart, D. L. 1983. Selection in Chemostats. Microbiol. Rev. 47-2: 150-168.

11. El-Sheekh, M., Abomohra, A.E., Hanelt, D. 2013. Optimization of biomass and fatty acid productivity of Scenedesmus obliquus as a promising microalga for biodiesel production. World J. Microbiol. Biotechnol. 29: 915-922.

12. Eppley, R.W., Coatsworth, J.L. 1996. Culture of the marine phytoplankter. Dunaliella tertiolecta, with light-dark cycles. Arch. Microbiol. 55: 66-80.

13. Fabregas, j., Maseda, a., Dominguez, a., Otero, a. 2004. The cell composition of Nannochloropsis sp. Changes under different irradiances in semicontinuous culture. World J Microbiol. Biotechnol. 20: 31-35.

14. Fabregas, j., Herrero, c., Cabezas, b., Abalde, j. 1985. Mass culture and biochemical variability of the marine microalga Tetraselmis suecica with high nutrient concentrations. Aquacult. J. 49: 231-244.

15. Fabregas, J., Abalde, J., Herrero, C., Cabezas, B., Veiga, M. 1984. Growth of the marine microalga Tetraselmis suecica in batch cultures with different salinities and nutrient concentrations. Aquacult. J. 42: 207-215.

16. Fan, L., Zhang, Y., Cheng, L., Zhang, L., Tang, D., Chen, H. 2007. Optimization of carbon dioxide fixation by Chlorella vulgaris cultivated in a membrane photobioreactor. Chem. Eng. Technol. 30-8: 1094-1099.

17. Fernandez-Sevilla, J.M., Molina-Grima, E., Garcia-Camacho, F., Acien-Fernandez, F.G., Sanchez-Perez, J.A. 1998. Photolimitation and photoinhibition as factors

Page 95: Effects of Light, Co2 and Reactor Design on Growth of Algae

81 

 

determining optimal dilution rate to produce eicosapentaenoic acid from cultures of the microalga Isochrysis galbana. Appl. Microbiol. Biotechnol. 50: 199-205.

18. Fidalgo, J.P., Cid, A., Torres, E., Sukenik, A., Herrero, C. 1998. Effects of nitrogen source and growth phase on proximate biochemical composition, lipid classes and fatty acid profile of the marine microalga Isochrysis galbana. Aquacult. J. 166: 105-116.

19. Garcia-Malea Lopez, M.C., Del Rio Sanchez, E., Casas lopez, J,L., Acien Fernandez, F.G., Fernandez Sevilla, J.M., Rivas, J., Guerrero, M.G., Molina Grima, E. 2006. Comparative analysis of the outdoor culture of Haematococcus pluvialis in tubular and bubble column photobioreactors. J. Biotechnol. 123: 329-342.

20. Garcia-Sanchez, J.L., Sanchez-Perez, J.A., Garcia-Camacho, F., Fernandez-Sevilla, J.M., Molina-Grima, E. 1996. Optimization of light and temperature for growing Chlorella sp using response surface methodology. Biotechnol. Tech. 10: 329-334.

21. Hill, J., Nelson, E., Tildman, D., Polasky, S., Tiffany, D. 2006. Environmental, economic, and energetic cost and benefits of biodiesel and ethanol biofuels. Proc. Nat. Acad. Sci. USA. 103-30: 11206-11210.

22. Jeon, B.H., Choi, J.A., Kim, H.C., Hwang, J.H., Abou-Shanab, R.A., Dempsey, B.A., Regan, J.M., Kim, J.R. 2013. Ultrasonic disintegration of micoalgal biomass and consequent improvement of bioaccessibility/bioavailability in microbial fermentation. Biotechnol. Biofuels. 6:1-9.

23. Kim, J.P., Kang, C.D., Park, T.H., Kim, M.S., Sim, S.J. 2006. Enhanced hydrogen production by controlling light intensity in sulphur-deprived Chlamydomonas reinhardtii culture. Int. J. Hydrogen Energy. 31: 1585-1590.

24. Krichnavaruk, S., Powtongsook, S., Pavasant, P. 2007. Enhanced productivity of Chaetoceros calcitrans in airlift photobioreactors. Bioresour. Technol. 98: 2123-2130.

Page 96: Effects of Light, Co2 and Reactor Design on Growth of Algae

82 

 

25. Kruse, O., Rupprecht, J., Mussgnug, J. H., Dismukes, C., Hankamer, B. 2005. Photosynthesis: a blueprint for solar capture and biohydrogen production technologies. Photochem. Photobiol. Sci. 4: 957-969.

26. Li, Y., Yuan, Z., Mu, J., Chen, D., Feng, B. 2013. Proteomic analysis of lipid accumulation in Chlorella protothecoides cells by heterotrophic N deprivation coupling cultivation. 27: 4031-4040.

27. Levasseur, M., Thompson, P. A., Harrison, P. J. 1993. Physiological acclimation of marine phytoplankton to different nitrogen sources. J. Phycol. 29: 587-595.

28. Matthijs, H.C.P., Balke, H., Van Hes, U.M., Kroon, B.M.A., Mur, L.R., Binot, R.A. 1996. Application of light-emitting diode in bioreactors: flashing light effects and energy economy in algal culture (Chlorella pyrenoidosa). Biotechnol. Bioeng. 50: 98-107.

29. Merchuk, j.C., Ronen, M., Giris, S., Arad, S. 1998. Light/dark cycles in the red growth of the Microalga Porhyridium sp. Biotechnol. Bioeng. 59: 706-713.

30. Molina-Grima, E., Fernandez-Sevilla, J.M., Sanchez-Perez, J.A., Garcia-Camacho, F. 1996. A study of simultaneous photolimitation and photoinhibition in dense microalgal cultures taking into account incident and averaged irradiances. J. Biotechnol. 45: 59-69.

31. Nayak, M., Rath, S.S., Thirunavoukkarasu, M., Panda, P.K., Mishra, B.K., Mohanty, R.C. 2013. Maximizing biomass productivity and CO2 biofixation of microalga Scenedesmus sp by using sodium hydroxide. J. Microbiol. Biotechnol. 23-9: 1260-1268.

32. Ogbonna, J. C., Soejima, T., Ugwu, C. U., Tanaka, H. 2001. An integrated system of solar light, artificial and organic carbon supply for cyclic photoautotrophic-heterotrophic cultivation of photosynthetic cells under day-night cycles. Biotechnol. Lett. 23: 1401-1406.

Page 97: Effects of Light, Co2 and Reactor Design on Growth of Algae

83 

 

33. Ogbonna, J.C., Soejima, T., Tanaka, H. 1999. An integrated solar and artificial light system for internal illumination of photobioreactors. J. Biotechnol. 70: 289-297.

34. Ogbonna, J.C., Tanaka, H. 1996. Night biomass loss and changes in biochemical composition of cells during light: dark cyclic culture of Chlorella pyrenoidosa. J. Ferment. Bioeng. 82: 558–564.

35. Park, S.J., Choi, Y.E., Kim, E.J., Park, W.K., Kim, C.W., Yang, J.W. 2012. Serial optimization of biomass production using microalga Nannochloris oculata and corresponding lipid biosynthesis. Bioprocess. Biosyst. Eng. 35: 3-9.

36. Pittman, J.K., Dean, A.P., Osukendo, O. 2011. The potential of sustainable algal biofuel production using wastewater resources. Bioresour. Technol. 102: 17-25.

37. Prommuak, C., Pavasant, P., Quitain, A.T., Goto, M., Shotipruk, A. 2013. Simultaneous production of biodiesel and free lutein from Chlorella vulgaris. 36-5: 733-739.

38. Quiang, H., Richmond, A. 1994. Optimizing the population density in Isochrysis galbana grown outdoors in a glass column photobioreactor. J. Appl. Phycol. 6: 391-396.

39. Spolaore, P., Claire, J.C., Elie, D. 2006. Commercial applications of microalgae. J. Bioscien. Bioeng. 101: 87-96.

40. Wang, C.Y., Fu, C.C., Liu, Y.C. 2007. Effects of using light-emitting diodes on the cultivation of Spirulina platensis. Biochem. Eng. J. 37: 21-25.

41. Yeh, N., Chung, J.P. 2009. High-brightness LEDs-energy efficient lighting sources and their potential in indoor plant cultivation. Renew. Sustain. Energ. Rev. 13: 2175-2180.

Page 98: Effects of Light, Co2 and Reactor Design on Growth of Algae

84 

 

42. Yun, Y.S., Park, J.M. 1997. Development of gas recycling photobioreactor system for microalgal carbon dioxide fixation. Korean J. Chem. Eng. 14-4: 297-1300.

Page 99: Effects of Light, Co2 and Reactor Design on Growth of Algae

85 

 

 

CHAPTER 5 BIOMASS PRODUCTION OF SIX ALGA SPECIES GROWING

UNDER LED AND SOLAR RADIATION

5.1 ABSTRACT

Three marine and three fresh water alga species were grown under two light treatments: a 24 h

continuous exposure to red, blue and red-blue LED illumination and a daily exposure to solar

radiation supplemented with nightly 12 h LED illumination. The algae were grown for 49 d

during which biomass measurements were taken every 7 d. The marine algae Nannochloris sp

produced an average biomass of 81-94 mg L-1 d-1; Tetraselmis sp produced 78-82 mg L-1 d-1 and

Dunnaliella salina produced 44-59 mg L-1 d-1. The fresh water alga Chlamydomonas sp produced

73-86 mg L-1 d-1; Chlorella vulgaris produced 60-86 mg L-1 d-1 and Scenedesmus sp produced

61-81 mg L-1 d-1. We found statistical differences in biomass production between algae species

and among light treatments (p = 0.0000), we also found statistical differences between marine

algae treatments (p = 0.0000) and between fresh water algae treatment (p = 0.0014).

5.2 INTRODUCTION

Systems for alga production have advanced during the past three decades from opens ponds to

closed tubular systems, to thin layer, airlift flat panel photo-bioreactors. Algal production in

closed reactors is 20 to 90% higher than in open systems. Thin layer, airlift flat panel photo-

bioreactors are more efficient in biomass production than open pond systems and tubular photo-

bioreactors because they provide better light utilization by reducing exposure periods of shadow

in the reactor. Also these reactors usually differ in media turbulence, velocity, oxygen saturation,

light source and intensity and duration of light exposure, including the flashing light effect

(Goldman, 1979).

Page 100: Effects of Light, Co2 and Reactor Design on Growth of Algae

86 

 

Light supply is an important factor controlling algal growth as it influences cell density and

biomass production. Open outdoor reactors have evolved in the past 30 y and are capable of

supporting biomass production rates of 15-25 g m-2 d-1 or 0.6-1 g L-1 d-1, with production peaks

of 30-40 g m-2 d-1 or 1.2-1.6 g L-1 d-1 (Goldman, 1979). Reports of commercial production of

Skeletonema sp in open systems have recorded biomass production rates ranging from 1.8 to 22.3

g m-2 d-1 or 0.072-0.89 g L-1 d-1, with peaks of 30 g m-2 d-1 or 1.2 g L-1 d-1 in June-August (Pauw

et al., 1983).

Development of closed-tubular bioreactor systems has evolved in Italy during the past 17 y,

doubling biomass production rates. Carlozzi (2008) reported a production increase from 25 g m-2

d-1 or 1.0 g L-1 d-1 in 1986 to 47.7 g m-2 d-1 or 1.9 g L-1 d-1 in 2003 for the algae Arthrospira

platensis and production rates of 0.49 g L-1 d-1 for Tetraselmis suecica. Comparable biomass

production rates for the algae Phaeodactylum tricornutum, 1.2 g L-1 d-1, were obtained using a

horizontal tubular air-lift photobioreactor exposed to solar radiation (Acien Fernandez et al.,

2001). Table 1 summarizes biomass production rates for different algal species.

Page 101: Effects of Light, Co2 and Reactor Design on Growth of Algae

87 

 

Table 5.1. Biomass production for 6 different alga species.

Algal specie  Biomass production  References 

Chlorella pyrenoidosa  2.5‐36 g m‐2d‐1 or 0.1‐1.44 g L‐1 d‐1  Ramos and Roux, 1986 

Chlorella vulgaris  13 mg L‐1 d‐1  Liang et al ., 2009 

Scenedesmus sp  26 mg L‐1 d‐1  Andrade et al., 2009 

Scenedesmus  dimorphus  1.5 mg L‐1 d‐1  Dev‐Goswami, 2011 

Scenedesmus quadricauda  1.3 mg L‐1 d‐1  Dev‐Goswami, 2011 

Scenedesmus almeriensis  870 mg L‐1 d‐1  Sanches et al., 2008 

Chlamydomonas reinhardtii  820‐2000 mg L‐1 d‐1  Kong et al., 2009 

Tetraselmis sp  1.35 g L‐1 d‐1  Garcia et al., 2012 

Tetraselmis suecica  1.07 g L‐1 5 d or 214 mg L‐1 d‐1  Go et al., 2012 

Tetraselmis suecica  64 mg L‐1 d‐1  Lee et al., 2011 

Tetraselmis sp  45 g m‐2 d‐1 or 1.8 g L‐1 d‐1  Huerlimann et al., 2010 

Tetraselmis suecica  7.6 g m‐2 d‐1 or 304 mg L‐1 d‐1  Bondioli et al., 2012 

Nannochloropsis oculata  20 mg L‐1 d‐1  Lee et al., 2011 

Nannochloropsis sp  13 g m‐2 d‐1 or 520 mg L‐1 d‐1  Huerlimann et al., 2010 

Nannochloropsis sp  9.9 g m‐2 d‐1 or 396 mg L‐1 d‐1  Bondioli et al., 2012 

Nannochloropsis oculata  0.58 g L‐1 5 d or 116 mg L‐1 d‐1  Su et al., 2011 

Nannochloris oculata  387‐1011 mg L‐1 d‐1  Park et al., 2012 

Dunaliella sp  25 g m‐2 d‐1 or 1 g L‐1 d‐1  Cortinas et al., 1984 

Dunnaliella salina  80 mg L‐1 d‐1  Garcia‐Gonzales et al., 2005 

Other systems for algal production include flat panel photo-reactors varying in reactor inclination

and depth, and airlift reactors. Production of Spirulina platensis was 0.3 g L-1 d-1 with a 10.4 cm

reactor depth, increasing to 4.3 g L-1 d-1 when the depth of the reactor decreased to 1.3 cm (Hu et

al., 1996). A vertical photo-bioreactor of 3 cm depth and containing baffles to provide a flashing

light exposure resulted in Chlorella vulgaris biomass production of 1.95 g L-1 d-1, with

production peaks of 2.6 g L-1 d-1 (Degen et al., 2001). Biomass production of Chlorella sp in an

outdoor, thin layer photo-bioreactor reached 30 g m-2 d-1 or 1.2 g L-1 d-1, the annual biomass

production estimated to be 80-100 tons ha-1 y-1 (Doucha and Livansky, 2009).

Page 102: Effects of Light, Co2 and Reactor Design on Growth of Algae

88 

 

The objective of this study was to evaluate biomass production of three marine and three fresh

water algal species grown under two light treatments: a 24 h continuous exposure to LED

illumination and a daily exposure to solar radiation supplemented with nightly 12 h LED

illumination.

5.3 MATERIALS AND METHODS

5.3.1 MICROORGANISM

Five of the six algal cultures used in this study were obtained from the algal culture collection

located at the University of Texas, Austin, TX. They included Chlorella vulgaris (26),

Nannochloris sp (1268), Scenedesmus sp (1589), Tetraselmis sp (2767), and Dunnaliella salina

(1644). A species of fresh water alga, Chlamydomonas sp, was isolated from a local liquid

manure tank (Fredonia, Colombia), and identified from photographic records.

Cultures were established in 1 L batch cultures for 15 d and then transferred to 4-L reactors. The

population density was established using direct microscopic counting techniques. The population

density at the start of the experiment was set at 4.5- 5*105 cells mL-1 by adjusting the volume of

medium in the reactors.

5.3.2 BIOREACTOR SYSTEM

Glass containers of 4 L capacity were filled with 3.6 L of nutrient media. Six replicate containers

were kept in a darkened room and provided continuous illumination from a LED source; 6

replicate containers were placed outdoors and received daily solar radiation and 12 h LED

illumination during the night.

5.3.3 LIGHT SOURCE

The LED lighting systems, contained in a sealed, transparent plexiglass container to protect

against water damage, were placed in the center of each bioreactor. The light emission intensity

Page 103: Effects of Light, Co2 and Reactor Design on Growth of Algae

89 

 

of each LED was 2143 µmol m-2 s-1 for blue LEDs and 1345 µmol m-2 s-1 for red LEDs. The

lighting system in each reactor contained 26 red LEDs and 26 blue LEDs for a total light power

of 4.94 W (2.34 W for red and 2.6 W for blue) or a power/bioreactor volume ratio of 1.37 W L-1.

Light emission intensity was measured in µmol m-2 s-1 using a quatum meter (Model MQ,

Apogee Instruments Inc., Logan, UT, USA).

5.3.4 LIGHT DISPERSION

LEDs of 0.09-0.1W can produce emission from 1345 to 2143 µmol m-2 S-1, this radiation is more

than the optimum radiation for photosynthesis activity; 50-250 µmol m-2 S-1 therefore the

radiation from the LED could cause photo-inhibition and photo-damage which leads to the

reduction of the photosynthetic activity (Kim et al., 2006; Degen et al.,2001; Ogbonna et al.,

1999; Merchuk et al., 1998; Fernandez-Sevilla et al., 1998; Molina –Grima et al., 1996, Garcia-

Sanchez et al., 1996; Qiang and Richmond, 1994; Aiba, 1982; Bannister, 1979; Eppley and

Coatsworth, 1966). The dispersion of light in algae media develops in short distances, in this

experimental setup the distance from the LED bulb to the algal in suspension is 7 mm. Enough

distance to reduce the light intensively from 2143 µmol m-2 s-1 to 300 µmol m-2 s-1 or even lower,

therefore avoiding photo inhibition process, however the distance is enough to support optimum

algal phototropic growth under artificial conditions (Ogbonna et al., 1999).

5.3.5 EXPERIMENT 1: 24 H RED + BLUE LEDS

The illumination was provided by LEDs (Alas, China) and consisted of red emitting diodes at

620-625 nm and blue at 425-430 nm. The illumination system was driven by a power supply of

12 V and 8 amp (Model SA 201-3485, ASTEC, California, USA).

5.3.6 EXPERIMENT 2: DAILY SOLAR + 12 H LEDS

Six reactors were exposed to daily solar radiation and provided red (620-625nm) and blue (425-

430nm) LED illumination for 12 h during the night.

Page 104: Effects of Light, Co2 and Reactor Design on Growth of Algae

90 

 

5.3.7 MIXING, QUANTIFYING CARBON SOURCE UNDER AUTOTROPHIC GROWTH

AND TEMPERATURE MEASUREMENTS

Two Resun turbines (Model AC-9904, Resun, China) were used to supply CO2 to the reactors

and for gentle mixing of the algal culture, one located inside and the other outside. Each turbine

provide a volume of 12 L min-1 in 21.6 L of algal suspension (air/media volume ratio was 556

mL of air L-1 min-1 growing media). The CO2 concentration supplied in the air to the two

experiments ranged from 0.0323% to 0.0428%, measured by CO2 meter (Model 7001, Telaire-

General Electric, California, USA).

The temperature (inside and outside) fluctuated from 23°C + 2 as measured by an Infrared

thermometer (Model Fluke 62, Fluke Corporation, Everett, Washington, USA).

5.3.8 CULTURE MEDIUM

Commercial fertilizer (Solucat 25-5-5, Atlantica Agricola, Villena, Spain) containing nitrogen

25% as ammonia 14.2% plus nitrates 10.8%. P2O5 5%, K2O 5%, Fe 0.020%, Mn 0.010%, Zn

0.002%, B 0.010%, Cu 0.002%. We used 0.8 g of fertilizer per litre of alga suspension for both

fresh water and marine alga, at the beginning of the experiment the electrical conductivity was

0.95 ms + 0.01 ms, measured by EC sensor (Model HI 991301, Hanna, Michigan, USA). The

three marine algae were amended with 3.5% of NaCl after the application of the solucat

fertilizer.

5.3.9 BIOMASS DETERMINATION

For biomass quantification we used the dry weight methodology which is divided in 7 steps:

1. Dry a 15 mL glass tubing until reach constant weight.

2. Fill out the tubing with 10mL of the algae media from the experiment.

Page 105: Effects of Light, Co2 and Reactor Design on Growth of Algae

91 

 

3. Centrifuge the tubing at 4000 rpm for 15 min.

4. Separate the water residues in the tubing.

5. Dry the tubing (105 oC) with the algae in an oven until reaches constant weight.

6. Weigh the tubing with the algae and subtract the weight of the tubing without the algae.

7. The resulting value is the alga biomass in a volume of 10mL. We repeated this procedure in

duplicate each time alga biomass was measured.

5.3.10 EXPERIMENTAL DESIGN AND STATISTICAL ANALYSIS

The experiment measured standing biomass for six algae species; three marine Tetraselmis sp,

Dunnaliella salina and Nannochloris sp and three fresh water Chlorella vulgaris, Scenedesmus

sp and Chlamydomonas sp. The algae were growing under two different light conditions;

Experiment 1; Algae growing under 24 h red and blue LEDs and experiment 2; algae growing

under daily solar radiation and 12 h red and blue LEDs. The algae were grown for 49 d, every 7

d the biomass was measured before and after harvesting half of the reactor volume (1.8 L per

reactor). The growing media was 0.8 g L-1 of Solucat fertilizer for both experiments. The light

power ratio was 1.37 W L-1 and the air ratio was 556 mL L-1 for both experiments. The biomass

results were analyzed using statistical methods such as standard deviation, coefficient of

variance, and one way ANOVA and T test using Biostatistics 1.0.

5.4 RESULTS

5.4.1 BIOMASS PRODUCTION FOR FRESH WATER ALGAE GROWING UNDER TWO

DIFFERENT LIGHT TREATMENTS

The biomass production for three fresh water algae during 49 days of incubation period ranged

from 320 to 1070 mg L-1 week-1 and the average was from 423 to 627 mg L-1 week-1. The

standing biomass ranged from 900 to 1180 mg L-1, thus the biomass recovery after harvest was 7

d, showing a doubling rate of 6-7 days, and the coefficient of variance of the biomass production

Page 106: Effects of Light, Co2 and Reactor Design on Growth of Algae

92 

 

fluctuated from 9.6% to 32%. For all algal cultures, the light treatment consisting of daily solar

radiation and red + blue LEDs during the night resulted in more biomass production than

continuous illumination with red + blue LEDs. The cumulative biomass production during 49 d

is shown in Table 5.2. was for Chlorella sp 24 h LED 2960 mg L-1 in solar-LED 4230 mg L-1,

for Scenedesmus sp in 24 LED 3030 mg L-1, in solar-LED 3990 mg L-1, for Chlamydomonas sp

in 24h LED 3590 mg L-1, in solar-LED 4390 mg L-1. See Table 5.2 for details.

The experiment conducted on the 6 algae developed similar standing biomass; biomass

production ranged from 190 to 776 mg L-1 week-1, the average was from 317 to 659 mg L-1 and

the maximun standing biomass 1310 mg L-1. The alga Nannochloris sp developed the highest

biomass production 4614 mg L-1 under the light treatment solar-LED and 3980 mg L-1 under

LED. Tetraselmis sp developed more biomass under LED 4016 mg L-1 and 3840 mg L-1 under

solar-LED. The lowest biomass production recorded of all 6 algae was Dunnaliella salina 2220

mg L-1 under LED and 2870 mg L-1 under solar-LED. The growth of the marine algae is more

stable than the growth of the fresh water algae, therefore the coefficient of variance 7.6% to 22%

is lower than the coefficient of variance of the fresh water algae.

The algae Chlamydomonas sp developed the highest biomass production for the fresh water

algae, whereas Nannochloris sp produced more biomass than the other two marine algae and the

three fresh water algae including Chlamydomonas sp. The marine algae Nannochloris sp

produced highest biomass of all six algae grown in the present study. See Table 5.3 for more

information. 

 

Page 107: Effects of Light, Co2 and Reactor Design on Growth of Algae

93 

 

Table 5.2. Biomass production for three fresh water algae under two light sorces.

LEDs 24h Daily Solar + 12h LEDs LEDs 24h Daily  Solar + 12h LEDs LEDs 24h Daily Solar + 12h LEDs

Biomass mg L‐1 

Biomass mg L‐1 

Biomass mg L‐1 

Biomass mg L‐1 

Biomass mg L‐2

Biomass mg L‐3

0‐7 (1‐7) 480 490 430 530 500 1070

7‐‐15 380 690 420 570 600 590

15‐‐21 500 560 480 510 520 490

21‐‐28 370 710 500 630 470 630

28‐‐35 420 600 400 580 380 530

35‐‐42 470 590 480 650 570 610

42‐‐49 340 590 320 520 550 470

Total 2960 4230 3030 3990 3590 4390

Mean 422.9 604.3 432.9 570.0 512.9 627.1

Max 500.0 710.0 500.0 650.0 600.0 1070.0

Min 340.0 490.0 320.0 510.0 380.0 470.0Standard Deviation 61.8 75.2 61.8 54.5 73.0 204.4

Coefficient of Variation 0.146 0.125 0.143 0.096 0.142 0.326

Max standing biomass biomass mg L‐1  1010 1260 920 1180 1080 1180

Mean algal biomass production for 3 fresh water algae during 49 days of incubationChlamydomonas sp

Incubation days

Chlorella vulgaris Scenedesmus sp

Table 5.3. Biomass production for three marine algae under two light sources.

LEDs 24h Daily Solar + 12h LEDs LEDs 24h Daily Solar + 12h LEDs LEDs 24h Daily Solar + 12h LEDs

Biomass mg L‐1 

Biomass mg L‐1 

Biomass mg L‐1 

Biomass mg L‐1 

Biomass mg L‐2

Biomass mg L‐3

0‐7 (1‐7) 560 520 590 770 190 450

7‐‐15 350 550 570 630 310 460

15‐‐21 776 540 560 544 400 400

21‐‐28 590 490 480 720 350 400

28‐‐35 600 540 620 670 290 430

35‐‐42 550 580 610 620 350 300

42‐‐49 590 620 550 660 330 430

Total 4016 3840 3980 4614 2220 2870

Mean 573.7 548.6 568.6 659.1 317.1 410.0

Max 776.0 620.0 620.0 770.0 400.0 460.0

Min 350.0 490.0 480.0 544.0 190.0 300.0Standard Deviation 124.5 41.8 46.7 72.7 66.0 53.5

Coefficient of Variation 0.217 0.076 0.082 0.110 0.208 0.131

Max standing biomass biomass mg L‐1  1230 1220 1190 1310 970 1050

Mean algal biomass production for 3 marine algae during 49 days of incubation

Incubation days

Tetraselmis sp Nannochloris sp Dunnaliella salina

Page 108: Effects of Light, Co2 and Reactor Design on Growth of Algae

94 

 

 

5.4.2 MEAN AND TOTAL BIOMASS PRODUCTION

This study shows higher biomass production in five of the six algae species grown under daily

solar radiation + 12 h LED illumination. Three of the algae, a marine and two fresh water,

produced a weekly mean of more than 600 mg L-1of biomass and a cumulative biomass

production reached 4 g L-1 during the 49 d of incubation (Figure 5.1 and 5.2).

Figure 5.1. Mean weekly algal biomass production under two light treatments during 49 d of incubation.

The higest weekly mean biomass production in 49 days was 659 mg L-1 obtained by the algal

Nannochloris sp under the light treatment; daily solar plus 12 h LED during night periods. The

ANOVA test for the entire experiment six algae species growing under two light treatment with

7 repetitions of seven days resulted in F = 9.3594 with p = 0.0000, this test prove that there is

Page 109: Effects of Light, Co2 and Reactor Design on Growth of Algae

95 

 

differences among algae treatment and light treatment. Figure 5.2 shows total biomass

production during the 49 days experiment.

Figure 5.2. Total algal biomass production under two light treatments.

ANOVA was performed to compare biomass production of algal species growing under the two

light treatments. The results for the three fresh water algae F = 5.0021 with p= 0.0014 and the

ANOVA for the three marine algae F = 20.6136 with p = 0.0000. These ANOVA tests show that

there is a statistically significant difference among light treatments and algae species with p <

0.01. We also performed an ANOVA for each alga to evaluate the different light treatments.

Table 5.4 shows statistical differences in light treatment for the algae Chlorella sp, Scenedesmus

sp with p < 0.01 and for Nannochloris sp and Dunnaliella salina with p < 0.02.

Page 110: Effects of Light, Co2 and Reactor Design on Growth of Algae

96 

 

Table 5.4. ANOVA test for total biomass production.

Chlorella vulgaris Scenedesmus sp Chlamydomonas sp Tetraselmis sp Nannochloris sp Dunnaliella salinaF = 24.2907 F = 19.3885 F = 1.9406 F = 0.2567 F = 7.6842 F = 8.3553

P = 0.0003 p = 0.0009 p = 0.1889 p = 0.6216 p = 0.0169 p = 0.0136

Algae ANOVA light treatments ( 24 h LEDs against daily solar plus 12 h LEDs in six algae)

Note: Statistical differences in bold p < 0.05

We performed 60 T tests to identify statistical differences between each individual algae and for

each light treatment. The T test for the light treatment 24 h LED showed differences between

algae resulting in statistical difference with p < 0.05 in 11 of the 15 test. The only cases where

there were no differences were for Chlorella sp against Scenedesmus sp, Chlamydomonas sp

against Scenedesmus sp, Chlamydomonas sp against Nannochloris sp and Tetraselmis sp against

Nannochloris sp. For the treatment of daily solar radiation plus 12 h LED and comparing algae

against algae, we found statistical differences in 7 of the 15 tests. For the T test in algae with 24

h LEDs against algae with daily solar plus 12 h LED, 11 out of the 15 tests showed statistical

differences with p < 0.05. For the test for algae with the light treatment daily solar plus 12 h LED

illumination against algae provided 24 h LED showed 7 out of the 15 with statistical difference.

In total 36 of the 60 individual tests comparing algal against algal resulted in statistical

difference with p < 0.05.

5.4.3 BIOMASS VARIANCE

The weekly mean production of marine algal biomass showed a lower coefficient of variance

than that of the fresh water algae, though the variance was lower in the fresh water algae growing

under continuous LED illumination. The marine algae developed lower average biomass

variance under daily solar radiation plus 12 h LED. Chlamydomonas sp and Dunnaliella salina

have the highest coefficient of variance of the algae species studied (Figure 5.3).

Page 111: Effects of Light, Co2 and Reactor Design on Growth of Algae

97 

 

Figure 5.3. Biomass variance during the experiment period.

5.5. DISCUSSION AND CONCLUSIONS

The light treatment combining daily solar radiation plus LED during the night produced greater

standing biomass and greater daily biomass production among five of the six algae species

during the 49 d experiment. There were statistical differences in biomass production for the two

light treatments among four of the six algae. The combination of solar radiation plus LED

illumination during the night periods produced more biomass than continuous LED illumination.

There were also differences in biomass production among algae species; Nannochloris sp under

solar-LED produced 4614 mg L-1 with an average daily production of 94 mg L-1 d-1; under

Page 112: Effects of Light, Co2 and Reactor Design on Growth of Algae

98 

 

continuous LED illumination biomass production was 3980 mg L-1 and daily average of 81 mg

L-1d-1. The second largest production was 4390 mg L-1 by Chlamydomonas sp under combined

solar-LED illumination and 3590 mg L-1 under continuous LED. Chlorella vulgaris produced

4230 mg L-1 under solar-LED and 2960 mgL-1 under continuous LED. Scenedesmus sp

developed 3990 mgL-1 under solar-LED and 3030 mgL-1 under continuous LED. The marine

alga Tetraselmis sp adapted well to the artificial illumination producing 4016 mg L-1 under

continuous LED exposure and 3840 mg L-1 under solar-LED. The lowest biomass production

was registered in Dunnaliella salina, 2870 mg L-1 under solar-LED illumination and 2200 mg L-1

under continuous LED. Statistical analysis shows significant differences with P >0.001 in

biomass production among fresh water algae, marine algae, and the 6 algae species together.

The mean daily biomass production by Nannochloris sp under the present study 81-94 mg L-1 d-1

is higher than the 20 mg L-1 d-1 reported by Lee et al., (2011) but is lower than that reported by

Bondioli et al. (2012) of 396 mg L-1 d-1 , 1011 mg L-1 d-1 by Park et al. (2012); 116 mg L-1 d-1 by

Su et al. (2011) and 520 mg L-1 d-1 by Huerlimann et al. (2010). The daily biomass production by

Chlamydomonas sp 73-86 mg L-1 d-1 is much lower than the 820-2000 mg L-1 d-1 reported by

Kong et al. (2009). The mean biomass production of Chlorella vulgaris at 60-86 mg L-1 d-1 is

similar to 70 mg L-1 d-1 reported by Zhao et al. (2013), but higher than 50 mg L-1 d-1 reported by

Kong et al. (2011), also higher than 13 mg L-1 d-1 reported by Liang et al. (2009) and lower than

100-1440 mg L-1 d-1 reported by Ramos and Roux (1986). The alga specie Scenedesmus sp

produced a mean biomass of 61-81 mg L-1 d-1 which is higher than 1.3-1.5 and 26 mg L-1 d-1

reported by Goswami (2011) and Andrade et al. (2009), but lower than 870 mg L-1 d-1 reported

by Sanches et al., 2008). The marine alga Tetraselmis sp developed a mean daily biomass of 78-

82 mg L-1 d-1 which was higher than 64 mg L-1 d-1 reported by Lee et al. (2011) and much lower

than that reported by Bondioli et al. (2012) at 304 mg L-1 d-1, 1350 mg L-1 d-1 by Garcia et al.

(2012), 214 mg L-1 d-1 by Go et al. (2012) and 1800 mg L-1 d-1 by Huerlimann et al. (2010). The

alga Dunnaliella salina grew in a slower rate than the other algal species in this study. Its mean

biomass production was 44-59 mg L-1 d-1 which was lower than results previously reported at 80-

1000 mg L-1 d-1 by Garcia-Gonzales et al. (2005) and Cortinas et al. (1984) (Table 5.1).

Page 113: Effects of Light, Co2 and Reactor Design on Growth of Algae

99 

 

The marine algae Tetraselmis sp was better adapted for growth under artificial illumination

producing greater biomass under continuous LED exposure than under the daily solar-LED

combination. Biomass production by Tetraselmis sp under continuous LED illumination was the

highest among the 6 algal species, though it was among the lowest under solar-LED illumination.

The coefficient of variance ranged from 7 to 32%; the fresh water algae Chlamydomonas sp

developed the highest coefficient of variance of the 6 algae species under solar-LED the CV was

32% and 14.2 % under continuous LED. The marine algae Tetraselmis sp had a CV 21.7% under

continuous LED and 7.6% under solar-LED.

From this research we can conclude with statistical support that growing algae with illumination

that combines both solar radiation and artificial LED lighting during the night results in greater

biomass production than continuous artificial illumination alone. This is because algae has

evolved with solar illumination.

Page 114: Effects of Light, Co2 and Reactor Design on Growth of Algae

100 

 

5.6 REFERENCES

1. Acien-Fernandez, F. G., Fernandez-Sevilla, J. M., Sanchez-Perez, J. A., Molina-Grima, E., Chisti, Y. 2001. Airlift-driven external-loop tubular photobioreactors for outdoor production of microalgae: assessment of design and performance. Chem. Eng. Sci. 56: 2721-2732.

2. Aiba, S. 1992. Growth kinetics of photosynthetic microorganisms. Adv. Biochem. Eng. j, 23: 85-156.

3. Andrade, R. C. E., Vera, B. A. L., Cardenas, L. C. H., Morales, A. E. D. 2009. Biomass production of microalga Scenedesmus sp with wastewater from fishery. Rev. Tec. Ing. Univ. zulia. 32-2: 126-134.

4. Bannister, T.T. 1979. Quantitative description of steady state, nutrient-saturated algal growth, including adaptation. Limnol. Ocean. 24: 76-96.

5. Biondioli, P., Bella, L. D., Rivolta, G., Zittelli, G. C., Bassi, N., Rodolfi, L., Casini, D., Prussi, M., Chiaramonti, D., Tredici, M. R. 2012. Oil production by the marine microalgae Nannochloropsis sp. F&M-M24 and Tetraselmis suecica F&M-M33. Bioresour. Technol. 114: 567-572.

6. Carlozzi, P. 2008. Closed photobioreactor assessments to grow, intensively, light dependent microorganisms: a twenty-year Italian outdoor investigation. The Open Biotechnol J. 2: 63-72.

7. Cortinas, T. I., Silva, H. J., Ertola, R. j. 1984. Photosynthetic production of Dunnaliella biomass from natural salt water and carbon dioxide. J. Chem. Tech. Biotechnol. 34B: 291-295.

8. Degen, J., Uebele, A., Retze, A., Schmid-Staiger, U., Trosch, W. 2001. A novel airlift photobioreactor with baffles for improved light utilization through the flashing light effect. J. Biotechnol. 92: 89-94.

Page 115: Effects of Light, Co2 and Reactor Design on Growth of Algae

101 

 

9. Dev-Goswami, R. C. 2011. Scenedesmus dimorphus and Scenedesmus quadricauda two potent indigenous microalgae strains for biomass production and CO2 mitigation – a study on their growth behavior and lipid productivity under different concentration of urea as nitrogen source. J. Algal Biomass Utln. 2-4: 42-49.

10. Doucha, J., Livansky, K. 2009. Outdoor open thin-layer microalgal photobioreactor: potential productivity. J. Appl. Phycol. 21: 111-117.

11. Eppley, R.W., Coatsworth, J.L. 1996. Culture of the marine phytoplankter. Dunaliella tertiolecta, with light-dark cycles. Arch. Microbiol. 55: 66-80.

12. Fernandez-Sevilla, J.M., Molina-Grima, E., Garcia-Camacho, F., Acien-Fernandez, F.G., Sanchez-Perez, J.A. 1998. Photolimitation and photoinhibition as factors determining optimal dilution rate to produce eicosapentaenoic acid from cultures of the microalga Isochrysis galbana. Appl. Microbiol. Biotech. 50: 199-205.

13. Garcia, L., Alvarez, X., Bravo, K., Peralta, J., Barriga, A. 2012. Obtaining microalgae biomass in a small scale reactor. Eur. Associ. Develop. Renew. Energ. Environ. Power Quality. 1-3.

14. Garcia-Gonzales, M., Moreno, J., Manzano, J. C., Florencio, F. J., Guerrero, M. G. 2005. Production of Dunaliella salina biomass rich in 9-cis-β-carotene and lutein in a closed tubular photobioreactor. J Biotechnol. 115: 81-90.

15. Garcia-Sanchez, J.L., Sanchez-Perez, J.A., Garcia-Camacho, F., Fernandez-Sevilla, J.M., Molina-Grima, E. 1996. Optimization of light and temperature for growing Chlorella sp using response surface methodology. Biotechnol. Tech. 10: 329-334.

16. Go, S., Lee, S. J., Jeong, G. T. 2012. Factors affecting the growth and the oil accumulation of marine microalgae, Tetraselmis suecica. Bioprocess. Biosyst. Eng. 35: 145-150.

Page 116: Effects of Light, Co2 and Reactor Design on Growth of Algae

102 

 

17. Goldman, J. C. 1979. Outdor algal mass cultures I applications. Water Res. 13: 1-19.

18. Hu, Q., Guterman, H., Richmond, A. 1996. A flat inclined modular photobioreactor for outdoor mass cultivation of photoautotrophs. Biotechnol. Bioeng. 51: 51-60.

19. Huerlimann, R., Nys, R., Heimann, K. 2010. Growth, lipid content, productivity, and fatty acid composition of tropical microalgae for scale-up production. Biotechnol. Bioeng. 107-2: 245-257.

20. Kim, J.P., Kang, C.D., Park, T.H., Kim, M.S., Sim, S.J. 2006. Enhanced hydrogen production by controlling light intensity in sulphur-deprived Chlamydomonas reinhardtii culture. Int. J. Hydrogen Energy. 31: 1585-1590.

21. Kong, W., Song, H., Cao, Y., Yang, H., Hua, S., Xia, C. 2011. The characteristics of biomass production, lipid accumulation and chlorophyll biosynthesis of Chlorella vulgaris under mixotrophic cultivation. Afr. J. Biotechnol. 10-55: 11620-11630.

22. Kong, Q. X., Li, L., Martinez, B., Chen, P., Ruan, R. 2010. Culture of microalgae Chlamydomonas reinhardtii in wastewater for biomass feedstock production. Appl. Biochem. Biotechnol. 160: 9-18.

23. Lee, S. J., Go, S., Jeong, G, T., Kim, S, K. 2011. Oil production from five marine microalgae for the production of biodiesel. Biotechnol. Bioprocess Eng. 16: 561-566.

24. Liang, Y., Sarkany, N., Cui, Yi. 2009. Biomass and lipid productivities of Chlorella vulgaris under autotrophic, heterotrophic and mixotrophic growth conditions. Biotechnol Lett. 31: 1043-1049.

25. Mandalam, R. K., Palson, B. O. 1998. Elemental balancing of biomass and medium composition enhances growth capacity in high density Chlorella vulgaris cultures. Biotechnol. Bioeng. 59-5: 605-611.

Page 117: Effects of Light, Co2 and Reactor Design on Growth of Algae

103 

 

26. Merchuk, J. C., Ronen, M., Giris, S., Arad, S. M. 1998. Light/dark cycles in the growth of the red microalga Porphyridium sp. Biotechnol. Bioeng. 59-6: 705-713.

27. Molina-Grima, E., Fernandez-Sevilla, J. M., Sanchez-Perez, J. A., Garcia-Camacho, F. 1996. A study on simultaneous photolimitation and photoinhibition in dense microalgal cultures takin into account incident and averaged irradiances. J. Biotechnol. 45: 59-69.

28. Ogbonna, J.C., Soejima, T., Tanaka, H. 1999. An integrated solar and artificial light system for internal illumination of photobioreactors. J. Biotechnol. 70: 289-297.

29. Park, S. J., Choi, Y. E., Kim, E. J., Park, W. K., Kim, C. W., Yang, J. W. 2012. Serial optimization of biomass production using microalga Nannochloris oculata and corresponding lipid biosynthesis. Bioprocess. Biosyst. Eng. 35: 3-9.

30. Pauw, N. D., Verboven, J., Claus, C. 1983. Large-scale microalgae production for nursery rearing of marine bivalves. Aquacult. Eng. 2: 27-47.

31. Pirt, S. J., Lee, Y. K., Walach, M. R., Pirt, M. W., Balyuzi, H. H., Bazin, M. j. 1983. A tubular bioreactor for photosynthetic production of biomass from carbon dioxide design and performance. J. Chem. Tech. Biotechnol. 33B: 35-58.

32. Quiang, H., Richmond, A. 1994. Optimizing the population density in Isochrysis galbana grown outdoors in a glass column photobioreactor. J. Appl. Phycol. 6: 391-396.

33. Ramos-de-Ortega, A., Roux, J. C. 1986. Production of Chlorella biomass in different types of flat bioreactors in temperate zones. Biomass. 10: 141-156.

34. Sanchez, J. F., Fernandez-Sevilla, J. M., Acien, F. G., Ceron, M. C., Perez-Parra, J., Molina-Grima, E. 2008. Biomass and lutein productivity of Scenedesmus almeriensis: influence of irradiance, dilution rate and temperature. Appl. Microbiol. Biotechnol. 79: 719-729.

Page 118: Effects of Light, Co2 and Reactor Design on Growth of Algae

104 

 

35. Su, C. H., Chien, L. J., Gomes, J., Lin, Y. S., Yu, Y. K., Liou, J. S., Syu, R. J. 2010. Factors affecting lipid accumulation by Nannochloropsis oculata in a two-stage cultivation process. J. Appl. Phycol. 23: 903-908.

36. Zhao, Y., Wang, J., Zhang, H., Yan, C., Zhang, Y. 2013. Effects of various LED light wavelengths and intensities on microalgae-based simultaneous biogas upgrading and digestate nutrient reduction process. Bioresour. Technol. 136: 461-468.

Page 119: Effects of Light, Co2 and Reactor Design on Growth of Algae

105 

 

 

CHAPTER 6 EFFECT OF USING DIFFERENT LEDS LIGHTS ON THE

GROWTH AND BIOMASS PRODUCTION OF CHLORELLA VULGARIS

 

6.1 ABSTRACT

Use of artificial light for indoor algae production is expensive so it is critical that the source of

the light energy be efficient for cost-effectiveness. The effects of continuous exposure to four

different LED illumination systems on the growth of Chlorella vulgaris were studied during a 15

d incubation. The light sources evaluated supplied white (380-760 nm), blue (430 nm), red (625

nm) and a combination red+blue wavelengths. The growth parameters measured included cell

density and biomass production. Exposure to white and to red+blue lights resulted in the highest

growth rates, reaching a population density of 1.8*107 cells mL-1, with a standing biomass of 1 g

L-1 and a maximum doubling rate of 1.1 d-1. Red light exposure resulted in the lowest cell density

of 1.04*107 cells mL-1 with a standing biomass of 0.57g L-1 and maximum doubling rate of 0.72

d-1. This study showed that light sources combining wavelengths absorbed by the photosynthetic

pigments of C. vulgaris have potential to maximize growth.

Key words: Artificial light; illumination system; algal production; light wavelength absorption

6.2 INTRODUCTION

Alga biomass has become an important source of high-valued products for a wide variety of uses

(Hejazi and Wijffels, 2004; Sanches et al., 2002; Lorenz and Cysewski, 2000). Their products

include carotenoids and β-carotene photosynthetic pigments, astaxanthins, cantaxanthin, lutein,

phycocyanin, which can be used as natural colorants and antioxidants (Garcia-Gonzalez et al.,

2005; Del campo et al., 2004; Lorenz and Cysewski, 2000; Vandamme, 1994), proteins for

Page 120: Effects of Light, Co2 and Reactor Design on Growth of Algae

106 

 

human and animal feed stocks and enzymatic reactions, and therapeutics for a source of

renewable pharmaceuticals (Borowitzka, 1995; Schwartz et al., 1995). Some of the products are

antibacterial, antiviral, antitumor, and antifungal (Cannell, 1993). Algae also produce

polyunsaturated fatty acids rich in Omega 3, 6 and 9 and eicosapentanoic acid (Viswanath et al.,

2010; Belarbi et al., 2000) that can be used for nutritional application and biofuel production.

The alga Chlorella sp have been used in the pharmaceutical industry as a source of

carbohydrate extract for the immune system in the production of anti-flu (Pulz and Gross, 2004).

Studies have reported the use of Chlorella sp extracts for their antibiotic properties against gram

positive and gram negative bacteria, and for the growth inhibition of four cancer cell lines

(Ordog et al., 2004; Metting and Pyne, 1986). Chlorella sp have also been used in preparations

of health food supplements and as a source of amino acids (Benemann et al., 1987), Global

production of this species at 2000 tons year-1 is second largest for single algal cell culture (Pulz

and Gross, 2004).

Light energy has a central role for controlling algal growth under a phototropic regime

and, when these organisms are grown under artificial conditions, light becomes the most

expensive single input. Therefore, the quality of the light, the wavelength of the photon, as well

as photon intensity and duration (referred to as the flash effect or light/dark cycles), are

important for induction of phototropic growth and regulation of growth rate (Wang et al., 2007;

Katsuda et al., 2006; Jeon et al., 2005, 2006). Light emitting diodes (LEDs) can be used to

supply photons at the wavelengths that the algae absorb and at an appropriate intensity to support

phototropic growth (Yeh and Chung, 2009). Thus, LED technology is efficient at converting

electrical energy into light energy and this is an advantage for algae production (Michel and

Eisentraeger, 2004). This study examined the growth and biomass production of Chlorella

vulgaris under 4 LED light sources: red (625 nm), blue (430 nm), red+blue, and white (380-760

nm) over a 15-d incubation.

Page 121: Effects of Light, Co2 and Reactor Design on Growth of Algae

107 

 

6.3 MATERIALS AND METHODS

6.3.1 MICROORGANISM

Chlorella vulgaris (UTEX 26) was obtained from the algal culture collection located at the

University of Texas, Austin, TX. The culture was established in 1 L batch cultures for 15 d and

then transferred to 4-L reactors. The population density was determined using direct microscopic

counting techniques. The population density at the start of the experiment was set at 4.5- 5*105

cells mL-1 by adjusting the volume of medium in the reactors.

6.3.2 LIGHT SOURCE

Four light sources providing blue, red, white and red+blue illumination were evaluated. The light

sources were 0.1 W LEDs obtained from ALAS™ with the following emission parameters: red,

620-625 nm wavelength at 1345 µmol m-2 s-1 intensity; blue, 425-430 nm wavelength at 2143

µmol m-2 s-1; white, 380-760 nm wavelength at 1680 µmol m-2 s-1 intensity. Each light source

contained 52 LEDs fixed into a clear plexiglass tube (internal diameter, 4 cm) and placed in the

centre of the 4-L culture flask (internal diameter 12 cm). The illumination system was powered

by a transformer (Model SA 201-3485, ASTEC, California, USA) supplying 12 V and 8 amp. To

ensure that external illumination would not affect the experiment, the laboratory was kept dark

throughout the incubation period.

6.3.3 LIGHT EMISSION INTENSITY

Light intensity was measured using a quantum meter (Model MQ, Apogee Instruments Inc.,

Logan, UT, USA), and with a Lux Meter (Model MS6610, VIA Instruments, Shanghai, China).

Light emission from a LED is intense, a 0.1 W bulb produces from 1700 to 2300 µmol m-2 s-1

and this amount of energy greatly exceeds the maximum photosynthetic light absorption capacity

of algae, which is 50-250 µmol m-2 s-1. This could cause photo-inhibition and damage to the

photosynthetic system which would reduce the photosynthetic activity of the cell (Kim et al.,

Page 122: Effects of Light, Co2 and Reactor Design on Growth of Algae

108 

 

2006; Degen et al., 2001; Ogbonna et al., 1999; Merchuk et al., 1998; Fernandez-Sevilla et al.,

1998; Molina –Grima et al., 1996, Garcia-Sanchez et al., 1996; Qiang and Richmond, 1994;

Aiba, 1982; Bannister, 1979; Eppley and Coatsworth, 1966). Since the intensity of light

decreases as an inverse square of the distance from the source, the LEDs were positioned at a

distance from the algal suspension to provide an emission intensity level acceptable for

supporting photosynthetic growth (Table 6.1). In our study the distance from the LED source to

the algal suspension was 5 mm, a sufficient distance to reduce the light intensity to less than 300

µmol m-2 s-1, which is where the photo inhibition process begins, yet optimal for phototropic

growth (Ogbonna et al., 1999). Zhou and Richmond (1999) reported optimal growth of

Nannochloropsis sp in a vertical reactor located 10 cm from the light source.

Table 6.1. Effect of distance from the LED light source on light power.

Dispersion of light intensity from the source to the edges Light intensity (µmol m-2 s-1)

Distance from the light source (mm)

blue red white

0 2143 1345 1828 1 568 225 413 5 206 173 186 10 95 72 103 15 70 43 69 20 40 30 41 25 32 23 33 30 26 18 28 35 20 14 23 40 17 11 20 45 14 9 17 50 11 7 12

Page 123: Effects of Light, Co2 and Reactor Design on Growth of Algae

109 

 

6.3.4 REACTOR MANAGEMENT

Two air turbines (Model ACO-9720, Hailea, Guangdong, China) each with an output of 30 L

min-1, were used to mix the algal suspension. Carbon dioxide at a concentration of 3.7% ± 0.18%

was supplied at a rate of 864 mL min-1 L-1 of algal suspension from a compressed air tank

equipped with a mass flow regulator (Model 191 AR-60, Gentec Corporation, Shanghai, China).

The concentration of dissolved carbon dioxide in the algal suspension was measured at 5 s

intervals with a Telaire CO2 monitor (Model 7001). Room temperature 22.4°C + 0.5, was

measured with a CO2 temperature monitor (Model 7001, Telaire-General Electric, California,

USA); the algal suspension temperature was measured with a infrared thermometer (Model

Fluke 62, Fluke Corporation, Everett, Washington, USA).

6.3.5 CULTURE MEDIUM

The culture medium was prepared using a commercial synthetic fertilizer as the nutrient source,

Solucat (25-5-5 + trace elements), by dissolving 0.8 g in 1 L distilled water. At the start of each

experiment, the culture medium contained 114 mg L-1 ammonium-N, 86.4 mg L-1 nitrate-N, 40

mg L-1 phosphate, 40 mg L-1 potash, 160 μg L-1 iron, 80 μg L-1 manganese, 80 μg L-1 boron, 16

μg L-1 zinc, and 16 μg L-1 copper; the electrical conductivity was 0.95 ± 0.01 mS, measured by

EC meter (Model HI 991301, Hanna, Michigan, USA).

6.3.6 EXPERIMENTAL DESIGN AND STATISTICAL ANALYSIS

The experiment was set out in 24-4 L glass containers (internal diameter 12 cm), each containing

3 L of culture medium. Each light source was replicated 6 times and the illumination was on for

the entire time of the experiment. Algal cultures were kept at room temperature (~22ºC) and

incubated for 15 d. A one-way ANOVA and T test were conducted on the data using the

software Biostatistics 1.0 with α≤ 0.01. Figure 6.1, shows the physical set up of the experiment.

Page 124: Effects of Light, Co2 and Reactor Design on Growth of Algae

110 

 

Figure. 6.1. Experiment set up, including 24 replicates.

6.3.7 ALGAL POPULATION, LIGHT ABSORPTION PEAKS AND BIOMASS ANALYSIS

The population density was measured daily throughout the incubation by direct microscopy and

a cell-counting plate using the Neubauer chamber. The methodology is described below;

Shake each algae container to make sure homogenous algal distribution before taking a

20 mL sample with syringe

For very concentrated sample, dilute the sample with distillate water a ratio of 1:10

Fill the Neubauer chamber with algal media

Depending of the cell size, use a microscopic magnification of 10X or 40X

Count the cells in the chambers A and B, for the small cells (smaller than 6µm) count

those cells in the chamber C.

For cells in chamber A and B use this equation C = N * 104 * dil, where C = cells mL-1,

N = cells average in 1mm2 (1µL) and dil = dilution factor.

Page 125: Effects of Light, Co2 and Reactor Design on Growth of Algae

111 

 

For small cells use the results in chamber C and this equation; C = (N/4*10-6)*dil, where

C = cells mL-1, N = cell average in the 5 C chambers and dil = dilution factor.

 

Light absorption of each algal culture was measured over wavelengths ranging from 400 to 700

nm using the Shimadzu UV 1600 spectrophotometer. The colorimetric determination

methodology was used to calculate the standing biomass production at the end of the incubation.

Light absorption was measured using a spectrophotometer (Shimadzu UV 1600). A calibration

plot relating algal biomass with 7 biomass measurements from 0.25 g L-1 to 3.49 g L-1 against

light absorption measured at 680 nm resulted in linear equation Y = 0.7015X + 0.0362 with a R2

= 99.87%. The resulting equation was used to determine the standing biomass of the 24

replicates.

Page 126: Effects of Light, Co2 and Reactor Design on Growth of Algae

112 

 

 

6.4. RESULTS AND DISCUSSION

6.4.1 THE EFFECT LIGHT IN THE ALGAL ABSORPTION SPECTRA

The light absorbance pattern of an algal culture depends of the nature and quantity of the specific

photosynthetic pigments it produces to capture energy. Figure 6.2 shows the absorption spectra

of C. vulgaris grown under blue, red, red+blue and white light after 15 d incubation.

Figure. 6.2. Algal absorption spectra after 15-d incubation exposed to blue, red, red+blue and white LED lights.

The absorption spectrum of cultures exposed to white light can be used as a reference to evaluate

the effects of the other light sources on the formation of photosynthetic pigments. As shown in

Fig. 6.2, the alga contains photosynthetic pigments able to absorb light over the entire spectral

range, thus the main curvature is caused by the light absorption peaks of the main photosynthetic

pigments; chlorophyll a and b. Growth under white light showed the highest absorbance 1.55 +

0.019 with 1.52% coefficient of variance (CV), followed by that under red+blue 1.45 + 0.010

Page 127: Effects of Light, Co2 and Reactor Design on Growth of Algae

113 

 

with 0.83% CV; growth under blue light showed an absorbance of 1.15 + 0.016 with 1.59% CV

and that under red light 0.85 + 0.007 with 1.01% CV.

These data suggest that C. vulgaris growing under white light synthesized greater quantities of

photosynthetic pigments than when grown under light sources of restricted wavelengths.

The absorption spectrum of C. vulgaris does not appear to be altered by exposure to light of

different wavelengths. Since C. vulgaris is a fresh water species inhabiting shallow water bodies,

it may has evolved the ability to absorb solar visible radiation over a broad waveleght range.

Previous studies of the in vivo absorption spectra for Chlorella species, including Chlorella

vulgaris, have reported similar results. Minor differences in the blue and red absorption peaks

were observed in that the red absorption peak was higher than the blue peak (Kamiya, 1988;

Shioi and Sasa, 1984; Ley and Mauzerall, 1982; Shibata et al., 1954). The results of this study

suggest that C. vulgaris does not have a mechanism for altering it’s photosynthetic pigments to

be able to respond to radiation different from those wavelengths shown in Fig. 6.2.

6.4.2. POPULATION GROWTH OF C. VULGARIS EXPOSED TO DIFFERENT LIGHT SOURCES

Population growth under the four light sources shows a classical sigmoid curve (Fig. 6.3), with

slight differences in the adaptation time and period of exponential growth. The initial population

density was 4.5- 5*105 cells mL-1. Under white light, the growth rate during the exponential

growth phase was 0.67 d-1 and reached a maximum cell density of 1.80*107 cells mL-1. Under

blue light the growth rate was 0.64 d-1 with a maximum population density of 1.71*107 cells mL-

1. Growth under red light was the slowest with a growth rate at 0.43 d-1 and a population density

of 1.04*107 cells mL-1. Exposure to red + blue light resulted in a growth rate of 0.56 d-1 and a

maximum population density of 1.75*107 cells mL-1, which was not significantly different from

that under white light.

Page 128: Effects of Light, Co2 and Reactor Design on Growth of Algae

114 

 

Figure. 6.3. Growth curves of the four light treatments with 6 replicates of each treatment.  

The maximum population density obtained in this study is similar to that reported in previous

studies. Yi et al. (2000) obtained a cell density of 106 cells mL-1 and Mandalam and Palsson

(1995) reported cell densities ranging from 106 to 107 cells mL-1. Typical cell densities for

Chlorella sp under open pounds, bubble column reactors and static mixer reactors are in the

range of 3*104 to 1.38 *109 cells mL-1 (Cho et al., 2007; Borowitzka, 1999). Studies conducted

under heterotrophic growing conditions have obtained cell densities of 3.8*1008 cells mL-1

(Killam and Myers, 1956). Lee and Palsson (1996) reported cell densities ranging from 1*107 to

5*107 cells mL-1 in a growth study exposed to red (680 nm) light.

Page 129: Effects of Light, Co2 and Reactor Design on Growth of Algae

115 

 

6.4.3.3 STANDING BIOMASS PRODUCTION UNDER DIFFERENT LIGHT SOURCES

6.4.3.3.1 STANDING BIOMASS

The effects of light wavelength on biomass production are shown in Table 6.2. The highest

biomass production was 1044 mg L-1 in 15 d growing under white light followed by red + blue

at 1003 mg L-1, blue at 827 mg L-1 and least under red with a biomass of 571 mg L-1 . This

biomass is higher than 560 mg L-1 in red (660 nm) after 8 days incubation (Zhao et al., 2013), the

Zhao research also report biomass production in white 500 mg L-1, and in blue (460 nm) 370 mg

L-1, those values are lower than the biomass obtained in this study. Several studies of C. vulgaris

have shown that the typical standing biomass under both phototropic and heterotrophic growth

after 12-15 days of incubation is between 250 and 1700 mg L-1 (Bhola et al., 2010; Liang et al.,

2009); Chinnasamy et al., (2009) has reported a biomass of 210 mg L-1 growing in wastewater. A

most recent study has reached 2730 mg L-1 growing C. vulgaris under blue LED at 200 µmol m-2

s-1 and photoperiod of 12h light and 12h dark for 312 h (Atta et al., 2013).

Table 6.2. Standing biomass of C. vulgaris growing under white, blue, red and red + blue light.

Repetition Blue Red Red-Blue White1 813 568 987 10262 851 565 1007 10583 821 576 1006 10344 830 578 1005 10385 827 565 1001 10536 823 574 1014 1061

Mean 828 571 1003 1045Standard Deviation 13 6 9 14

Coefficient of Variation 0.016 0.010 0.009 0.014

Biomass production (mg L-1)

White light supported a greater standing biomass than the other light treatments because this

light source provided radiation over the entire range of photosynthetic pigments. Interestingly the

combination of red + blue light produced almost as much standing biomass as the white light.

Since this is the only treatment that combined more than one photosynthetic wavelength, C.

Page 130: Effects of Light, Co2 and Reactor Design on Growth of Algae

116 

 

vulgaris adapted better than the single wavelength treatment of either red or blue alone. Growth

of C. vulgaris under blue light was better than that under red because the absorption spectra has

one of the important peaks located in 430 nm and the other at 680 nm. From the growth and

biomass production we can propose that the algal do not appear to be able to change its

photosynthetic pigment composition to adapt to this specific wavelength.

In contrast to our study, Wang et al. (2007) reported that highest biomass production in Spirulina

platensis was achieved under red light (620-645 nm) followed by growth under white light (380-

760 nm), followed by green (515-540 nm), yellow (587-595 nm) and the least growth in blue

(460-475 nm). In our study C. vulgaris produced more biomass when expose to multiple

wavelengths such as the white light and the combination of red-blue, and a single wavelength

light not performing as well. The blue light (430 nm) is the best single wavelength light

treatment because it matches the wavelength absorption peak of the photosynthetic pigment of

this specie. Other studies in single wavelength light have been conducted to increase biomass

and astaxanthin concentration testing red and blue LEDs separately. The highest biomass

production was obtained using the blue LED 380-470 nm (Lababpour et al., 2005; Lababpour et

al., 2004). A study which compared blue 370-430 nm, red >580 nm and blue +red at an intensity

of 20-40 µmol m-2 s-1 for astaxanthin formation in the alga Haematococcus pluvialis, showed

that blue light and blue + red performed better than red light alone (Kobayashi et al., 1992).

This study analysed four LED light treatments with six repetitions per treatment. One way

ANOVA test was performed using the software Biostatistics 1.0. The ANOVA test resulted in F

= 2305 with p = 0.0000 showing statistical difference among the four treatments. Thus, the

exponential coefficient of variation was very low 1.27%. In addition, to complement the

Analysis of Variance (ANOVA test), we choose the student’s T test from the unequal variance

T-test, Mann-Whitney. We tested each treatment individually to one another using a Student’s t-

test at P ≤ 0.01, the reference value for the Student’s at p = 0.01 is T ≥ 4.6041. The results of

these analyses show that the biggest value is 100, made in red light against red-blue light, and the

Page 131: Effects of Light, Co2 and Reactor Design on Growth of Algae

117 

 

lowest value is 6.05 from the test made in red-blue light against white light. From the results of

the T test there is statistical difference at 99% confidence that each individual treatment differed

from each other.

6.4.3.4 EFFECT OF LIGHT SOURCE ON ALGAL CELL WEIGHT

Individual alga cell weight was calculated from the population cell density by dividing by the

standing biomass (Table 6.4). It ranged from ~49 to 64 pg (1 pg = 1*10-12 g) in cultures grown

under the four light sources. The heaviest cells were achieved under the white light at 63 pg cell-

1, followed by those grown under red + blue light at 61 pg cell-1, blue at 50 pg cell-1 and red at

55 pg cell-1. A study conducted by Krzywicka and Wagner (1975) reported cell weights ranging

from 7.0*10-11 to 8.2*10-11g cell-1. In a similar study, also using an inorganic fertilizer medium,

Griffiths (1963) reported cell weights ranging from 3.5*10-11 to 9.7*10-11g cell-1. Chimiklis and

Karlander (1973) recorded cell weights ranging 0.9*10-11 from 2.5*10-11 g cell-1 in a growth

study of Chlorella sorokiniana.

Table 6.4. Cell weight of C. vulgaris grown under 4 different light sources.

Repetition Red Blue Red + Blue White

1 5.33E‐11 4.90E‐11 5.97E‐11 6.22E‐11

2 5.43E‐11 5.11E‐11 6.22E‐11 6.43E‐11

3 5.62E‐11 4.93E‐11 6.13E‐11 6.27E‐11

4 5.62E‐11 4.99E‐11 6.07E‐11 6.18E‐11

5 5.54E‐11 4.98E‐11 6.03E‐11 6.25E‐11

6 5.54E‐11 4.89E‐11 6.02E‐11 6.39E‐11

Mean 5.51E‐11 4.97E‐11 6.07E‐11 6.29E‐11

Standard Deviation 1.11E‐12 8.28E‐13 8.99E‐13 9.95E‐13

Coefficient of Variation 0.0201 0.0167 0.0148 0.0158

Cell weight for 4 Light LEDs treatments

(Grams Cell‐1)

Page 132: Effects of Light, Co2 and Reactor Design on Growth of Algae

118 

 

The ANOVA test for cell weight was F = 227 with p = 0.0000 resulting in statistical difference,

the coefficient of variance was 1.69%. These tests suggests that the nature of the light source

had a significant influence on cell weight of Chlorella vulgaris.

6.5 CONCLUSIONS

Biomass production of the fresh water algal specie C. vulgaris was highest under white light,

presumably because this light source provided a larger wavelength photosynthetic spectra than a

single color light such as blue and red. The results of this study also show that C vulgaris cannot

change its photosynthetic machinery significantly to absorb light from other wavelengths

different from that of its evolutionary wavelength absorption spectra.

Biomass production under the four light treatments was significantly different from one another

(F> 99% of probability). Though the growth curves showed a similar pattern, the cell density and

growth rate was different for each individual light treatment, confirming the result of the biomass

production. Further, the different light sources significantly altered the cell weight of C. vulgaris.

Further studies of algal growth under red light (680 nm) are recommended to confirm that light

supplied at a single wavelength can develop more algal biomass than from multiple wavelength

light sources.

Page 133: Effects of Light, Co2 and Reactor Design on Growth of Algae

119 

 

6.6 REFERENCES

1. Aiba, S. 1982. Growth kinetics of photosynthetic microorganisms. Adv. Biochem. Eng. j. 23: 85-156.

2. Atta, M., Idris, A., Bukhari, A., Wahidin, S. 2013. Intensity of blue LED light: A potential for biomass and lipid content in fresh water microalgae Chlorella vulgaris. Bioresour. Technol. 148: 373-378.

3. Bannister, T.T. 1979. Quantitative description of steady state, nutrient-saturated algal growth, including adaptation. Limnol. Ocean. 24: 76-96.

4. Belarbi, E.H., Molina-Grima, E., Chisti, Y. 2000. A process for high yield and scaleable recovery of high purity eicosapentaenoic acid ester from microalgae and fish oil. Enz. Microb. Technol. 25: 516-529.

5. Benemann, J.R., Tillett, D., Weissman, J. 1987. Microalgae biotechnology. Trends. Biotechnol. 5: 47-53.

6. Bhola, V., Desikan, R., Santosh, S. K., Subburamu, K., Sanniyasi, E., Bux, F. 2011. Effects of parameters affecting biomass yield and thermal behaviour of Chlorella vulgaris. J. Bioscien. Bioeng. 111-3: 377-382.

7. Borowitzka, M.A. 1999. Commercial production of microalgae: ponds, tanks, tubes and fermenters. J. Biotechnol. 70: 313-321.

8. Borowitzka, M.A. 1995. Microalgae as sources of pharmaceuticals and other biologically active compounds. J. Appl. Phycol. 7: 3-15.

9. Cannell, R.J.P. 1993. Algae as a source of biologically active products. Pest. Scien. 39: 147-153.

Page 134: Effects of Light, Co2 and Reactor Design on Growth of Algae

120 

 

10. Chimiklis, P.E., Karlander, E.P. 1973. Light and calcium interactions in Chlorella inhibited by sodium chloride. Plant Physiol. 51: 48-56.

11. Chinnasamy, S., Ramakrishnan, B., Bhatnagar, A., Das, K. 2009. Biomass production potential of a wastewater alga Chlorella vulgaris ARC 1 under elevated levels of CO2 and temperature. Int. J. Mol. Scien. 10: 518-532.

12. Cho, S.H; Ji, S., Hur, B.S., Bae, J., Park, I., Song, Y. 2007. Optimun temperature and salinity conditions for growth of green algae Chlorella ellipsoidea and Nannochloris oculta. Fish. Scien. 73: 1050-1056.

13. Del Campo, J.A., Rodriguez, H., Moreno, J., Vargas, M.A., Rivas, J., Guerrero, M.G. 2004. Accumulation of astaxanthin and lutein in Chlorella zofingiensis (Chlorophyta). App Microb Cell Physiol. 64: 848-854.

14. Degen, J., Uebele, A., Retze, A., Schmid-Staiger, U., Trosch, W. 2001. A novel airlift photobioreactor with baffles for improved light utilization through the flashing light effect. J Biotechnol. 92: 89-94.

15. Eppley, R.W., Coatsworth, J.L. 1996. Culture of the marine phytoplankter. Dunaliella tertiolecta, with light-dark cycles. Arch. Microbiol. 55: 66-80.

16. Fernandez-Sevilla, J.M., Molina-Grima, E., Garcia-Camacho, F., Acien-Fernandez, F.G., Sanchez-Perez, J.A. 1998. Photolimitation and photoinhibition as factors determining optimal dilution rate to produce eicosapentaenoic acid from cultures of the microalga Isochrysis galbana. Appl. Microbiol. Biotechnol. 50: 199-205.

17. Garcia-Gonzalez, M., Moreno,J., Manzano, J.C., Florencio, F.J., Guerrero, M.G.2005. Production of Dunaliella salina biomass rich in 9-cis-β-carotene and lutein in a closed tubular photobioreactor. J. Biotechnol. 115: 81-90.

Page 135: Effects of Light, Co2 and Reactor Design on Growth of Algae

121 

 

18. Garcia-Sanchez, J.L., Sanchez-Perez, J.A., Garcia-Camacho, F., Fernandez-Sevilla, J.M., Molina-Grima, E. 1996. Optimization of light and temperature for growing Chlorella sp using response surface methodology. Biotechnol. Tech. 10: 329-334.

19. Griffiths, D.J. 1963. The effect of glucose on cell division in Chlorella vulgaris beijerinck (Emerson Strain). Ann Bot-London. 27-107: 493-504.

20. Hejazi, M.A., Wijffels, R.H. 2004. Milking of microalgae. Trends Biotechnol. 4: 189-194.

21. Jeon, Y.C., Cho, C.W., Yun, Y.S. 2006. Combined effects of light intensity and acetate

concentration on the growth of unicellular microalga Haematococcus pluvialis. Enzyme Microb. Technol. 39: 490-495.

22. Jeon, Y.C., Cho, C.W., YUN, Y.S. 2005. Measurement of microalgal photosynthetic activity depending on light intensity and quality. Biochem. Eng. J. 27: 127-131.

23. Kamiya, A. 1988. Blue light-induced in vivo absorbance changes and in vitro activation of nitrate reductase in a nitrate starved Chlorella mutant. Plant cell Physiol. 29-3: 489-496.

24. Katsuda, T., Shimahara, K., Shiraishi, H., Yamagami, K., Ranjbar, R., Katoh, S. 2006. Effect of Flashing light from blue light emitting diode on cell growth and astaxanthin production of Haematococcus pluvialis. J Bioscien. Bioeng. 102-5: 442-446.

25. Killam, A., Myers, J. 1956. A special effect of light on the growth of Chlorella vulgaris. Am. J. Botan. 43-8: 569-572.

26. Kim, J.P; Kang, C.D., Park, T.H., Kim, M.S., Sim, S.J. 2006. Enhanced hydrogen production by controlling light intensity in sulphur-deprived Chlamydomonas reinhardtii culture. Int. J. Hydrogen Energy. 31: 1585-1590.

Page 136: Effects of Light, Co2 and Reactor Design on Growth of Algae

122 

 

27. Kobayashi,M., Kakizono,T., Nishio, N., Nagai, S. 1992. Effects of light intensity, light quality, and illumination cycle on astaxanthin formation in a green alga, Haematococcus pluvialis. J Ferment. Bioeng. 74-1: 61-63.

28. Krzywicka, A. M., Wagner, G. H. 1975. Laboratory culture for C-labeling of Chlorella and Oedogonium. Int. J. Appl. Radiat. Is. 26: 515-518.

29. Lababpour, A., Shimashara, K., Hada, K., Kyoui, Y., Katsuda, T., Katoh, S. 2005. Fed-Batch culture under illumination with blue light emitting diodes (LEDs) for astaxanthin production by Haematococcus pluvialis. J. Bioscien. Bioeng. 100-3: 339-342.

30. Lababpour, A., Hada, K., Shimashara,K., Katsuda, T., Katoh, S. 2004. Effects of nutrients supply methods and illumination with blue light emitting diodes (LEDs) on Astaxanthin production by Haematococcus pluvialis. J. Bioscien. Bioeng. 98-6: 452-456.

31. Lee, C. G., Palsson, B. O. 1996. Photoacclimation of Chlorella vulgaris to red light from light emitting diodes leads to autospore release following each cellular division. Biotechnol. Progr. 12: 249-256.

32. Ley, A.C., Mauzerall, D.C. 1982. Absolute absorption cross-section for photosystem II and the minimum quantum requirement for photosynthesis in Chlorella vulgaris. Biochim. Biophys. Acta. 680: 95-106.

33. Liang, Y., Sarkany, N., Cui, Y. 2009. Biomass and lipid productivities of Chlorella vulgaris under autotrophic, heterotrophic and mixotrophic growth conditions. Biotechnol. Lett. 31: 1043-1049.

34. Lorenz, R.T., Cysewski, G.R. 2000. Commercial potential for Haematococcus microalgae as a natural source of astaxanthin. Trends. Biotechnol. 18: 160-167.

35. Mandalam, R.K., Palsson, B.O. 1995. Chlorella vulgaris (Chlorellaceae) does not secrete autoinhibitors at high cell densities. Am. J. Botany. 82-8: 955-963.

Page 137: Effects of Light, Co2 and Reactor Design on Growth of Algae

123 

 

36. Merchuk, J.C., Ronen, M., Giris, S., Arad, S. 1998. Light/dark cycles in the red growth of the microalga Porhyridium sp. Biotechnol. Bioeng. 59: 706-713.

37. Metting, B., Pyne, J. 1986. Biological active compounds from microalgae. Enzyme Microb. Technol. 8: 386-394.

38. Michel, K., Eisentraeger, A. 2004. Light-emitting diodes for the illumination of algae in ecotoxicity testing. Environ. Toxicol. 19: 609–613.

39. Molina-Grima, E., Fernandez-Sevilla, J.M., Sanchez-Perez, J.A., Garcia-Camacho, F. 1996. A study of simultaneous photolimitation and photoinhibition in dense microalgal cultures taking into account incident and averaged irradiances. J. Biotechnol. 45: 59-69.

40. Ogbonna, J.C., Soejima, T., Tanaka, H. 1999. An integrated solar and artificial light system for internal illumination of photobioreactors. J. Biotechnol. 70: 289-297.

41. Ordog, V., Stirk, W.A., Lenobel, R., Bancirova, M., Strnad, M., Van Staden, J., Szigeti, J., Nemeth, L. 2004. Screening microalgae for some potentially useful agricultural and pharmaceutical secondary metabolites. J. Appl. Phycol. 16: 309-314.

42. Pulz, O., Gross, W. 2004. Valuable products from biotechnology of microalgae. Appl. Microbial. Biotechnol. 65: 635-648.

43. Qiang, H., Richmond, A. 1994. Optimizing the population density in Isochrysis galbana grown outdoors in a glass column photo-bioreactors. J. Appl. phycol. 6: 391-396.

44. Sanches-Miron, A., Ceron-Garcia, M.C., Garcia-Camacho, F., Molina-Grima, E., Chisti, Y. 2002. Growth and biochemical characterization of microalgal biomass produced in bubble column and airlift photobioreactors studies in fed-batch culture. Enzyme Microb. Technol. 31: 1015-1023.

Page 138: Effects of Light, Co2 and Reactor Design on Growth of Algae

124 

 

45. Schwartz, R.E., Hirsch, C.F., Sesin, D.F., Flor, J.E., Chartrain, M., Fromtling, R.E., Harris, G.H., Salvatore, M.J., Liesch, J.M., Yudin, K. 1990. Pharmaceuticals from cultured algae. J. Ind. Microbiol. 5: 113-124.

46. Shibata, K., Benson, A.A., Calvin, M. 1954. The absorption spectra of suspensions of living microorganisms. Biochim. Biophys. Acta. 15: 461-469.

47. Shioi, Y., Sasa, T. 1984. Chlophyll formation in the YG-6 mutant of Chlorella regularis: Spectral characterization of protochlorophyllide phototransformation. Plant Cell Physiol. 25-1: 139-149.

48. Vandamme, E.J. 1994. The search for novel microbial fine chemicals, agrochemicals and biopharmaceuticals. J. Biotechnol. 37: 89-108.

49. Viswanath, B., Mutanda, T., White, S., Bux, F. 2010. The microalgae-a future source of biodiesel. Dynamic Biochemistry. Process Biotechnol. Mol. Biol. 4: 37-47.

50. Wang, C. Y., Fu, C.C., Liu, Y.C. 2007. Effects of using leght-emitting diodes on the cultivation of Spirulina platensis. Biochem Eng J. 37: 21-25.

51. Yeh, N., Chung, J.P. 2009. High-brightness LEDs-Energy efficient lighting sources and their potential in indoor plant cultivation. Renew. Sust. Energ. Rev. 13: 2175-2180.

52. Yi, L., Tahn, T.T., Zhong, C., Songsheng, Q., Ping, S. 2000. Microcalorimetric studies on the metabolism of Chlorella vulgaris. Chemosphere. 40: 845-849.

53. Zou, N., Richmond, A. 1999. Effect of light-path length in outdoor flat plate reactors on output rate of cell mass and of EPA in Nannochloropsis sp. J. Biotechnol. 70: 351-356.

54. Zhao, Y., Wang, J., Zhang, H., Yan, C., Zhang, Y. 2013. Effects of various LED light wavelengths and intensities on microalgae-based simultaneous biogas upgrading and digestate nutrient reduction process. Bioresour. Technol. 136: 461-468.

Page 139: Effects of Light, Co2 and Reactor Design on Growth of Algae

125 

 

 

CHAPTER 7 BIOMASS PRODUCTION OF MARINE AND FRESH WATER

ALGA GROWING UNDER FLASHING AND CONTINUOUS LED LIGHT

7.1 ABSTRACT

A marine alga Nannochloris sp and a fresh water alga Scenedesmus sp were grown under

flashing light conditions at 10Hz with a light dark ratio of 1:10. The illumination was prounded

by LED red 625 nm and blue 430 nm, the same algae were also grown under continuous light

using the same LED technology. The marine alga Nannochloris sp under flashing light reached a

cell density of 2.43*107 cell mL-1, developed a standing biomass of 1041 mg L-1 and the average

cell weight was 4.98*10-11 g cell-1. Under continuous light the cell density was 5.69*107 cells

mL-1, the standing biomass 892 mg L-1 and the cell weight was 2.26*10-11 g cell-1. In contrast the

fresh water alga Scenedesmus sp under flashing conditions developed a cell density of 1.65*107

cells mL-1, standing biomass of 1130 mg L-1 and an average cell weight of 8.50*10-11 g cell-1.

Under continuous light the cell density was 1.42*107 cells mL-1, the standing biomass 1018mg L-

1 and the average cell weight was 5.05*10-11 g cell-1. Both algae produced more final standing

biomass and more cell weight under flashing conditions, there is significant difference in daily

biomass production F = 81.7668 with P = 0.0000 between flashing and continuous light in

Nannochloris sp, however there is no significant difference in Scenedesmus sp.

7.2 INTRODUCTION

Continuous light may cause photo damage and photo inhibition which decreases photosynthetic

activity therefore flashing light is an alternative to reduce this damage (Fathurrahman et al.,

2013). The flashing effect increases algal biomass production, maximize energy efficiency

(Kong et al., 2013), and has been tested for different algal species. A study made in Chlorella sp

show an increase in cell density when the population was exposed to flashing light

Page 140: Effects of Light, Co2 and Reactor Design on Growth of Algae

126 

 

(Fathurrahman et al., 2013). Two independent studies using high frequency flashes of 0.5 to 1.0

Hz and 0.01 to 10 Hz with solar radiation and simulating solar radiation in Phaeodactylum

tricornutum produced an average of 3.7% more energy into biomass conversion over a 3 month

period and higher photosynthetic efficiency (Walsh and Legendre, 1988; Laws et al., 1983).

Similar studies simulating solar radiation with Chlorella sp have been made producing biomass

from 240 mg L-1 to 2720 mg L-1 (Pirt et al., 1980). More studies with the algal Chlorella kessleri

using flashing light effect of 10 KHz, 20 KHz and 50 KHz with light: darks of 1:9, 2:8, 3:7, 4:6

and 5:5 were conducted to determine the effect of flashing effect in cell volume (Park and Lee,

2001). The flashing effect was tested to analysed cell density; low and high frequency flashes

from 1Hz to 100Hz, and from 1KHz to 100 KHz with light period from 40 to 80% were tested

using the algal Isochrysis galbana, flashes of 10 KHz developed the highest cell density (Yago et

al., 2012).

The flashing effect can change algae pigment composition; flashes of 10 µS of light periods with

33-100 miliseconds of dark periods changed the chlorophyll concentration from 1 to 5% in the

algal Chlorella pyrenoidosa (Myers and Graham, 1971). Flashes of 1 to 100 Hz were tested

using the algal Chlamydomonas reinhardtii; flashes of 1 to 10Hz decreased the biomass

production in 10%, and cycles of 100Hz increased the biomass production in 35% compared to

continuous light (Vejraska et al., 2012). Thus flashes of 5 Hz to 37 KHz against continuous light

with the algal Chlorella kessleri were tested, higher cell density and oxygen production was

observed under flashes of more than 1KHz and 20% increase in cell density was achieved at 37

KHz (Park and Lee, 2000).

Medium frequency flashes have been studied too with the algae Chlamydomonas reinhardttii and

Chlorella sorokiniana, flashes of 13 to 87 s with light/dark fraction from 0.32-0.88 increases

chlorophyll content but decreased the biomass production of the algae. The results of the study

show that a dark period of less than 4.4 s between flashes does not affect biomass production

(Janssen et al., 1999). The algae Chlorella sp and Scenedesmus sp have been studied using

medium rate flashes of 1.2-260 s (Grobbelaar, 1991). Light periods of 0.52 s at 20 Wm-2 can

Page 141: Effects of Light, Co2 and Reactor Design on Growth of Algae

127 

 

support Chlorella sp at maximum photosynthetic activity for 9.2 s in the dark (Molina-Grima et

al., 1996; Pirt, 1986; Lee and Pirt, 1981), previous studies show that light periods of 1 ms at 24

Wm-2 support photosynthetic activity in Chlorella sp for 20 ms (Philiph and Myers, 1954).

Lower frequency flashes has been tested, 27, 60 and 110 s cycle, with the algal Porphyridium sp

using air lift and bubble column reactor concluded that a dark period of 6 s does not affect alga

performance (Merchuk et al., 1998). In conclusion flashing light may increase cell density, cell

volume, pigment composition, oxygen production, photosynthetic activity and biomass

production. The present investigation evaluated flashes of 10 Hz with light dark ratio of 1:10

using red and blue LEDs to measure cell density and biomass production against treatments with

continuous light using the marine algal Nannochloris sp and the fresh water algal Scenedesmus

sp.

7.3 MATERIALS AND METHODS

7.3.1 MICROORGANISM

The two alga species one marine and one fresh water used in this study are from the culture

collection of algae at the University of Texas at Austin UTEX; Nannochloris sp (1268) and

Scenedesmus sp (1589).

7.3.2 PHOTO REACTOR

The reactor had the following measurements: 30 cm diameter, and 110 cm length, it was filled

with the nutrient media to 80 cm of height with a active volume of 56 L. For the continuous light

experiment, we used a glass container of 4 L capacity filled with 3.6 L of nutrient media.

Page 142: Effects of Light, Co2 and Reactor Design on Growth of Algae

128 

 

7.3.3 LIGHT SOURCE AND FLASHING DEVICE

For the flashing light experiment a special computer device with a PLC was constructed to allow

control of 100 flashes per second, with different light dark ratios. This experiment was adjusted

to 10 flashes s-1 or 10 Hz with a light dark ratio of 1:10. In each flash cycle the lights were on for

10 millisec and off for 90 millisec for a total 100 millisec on and 900 millisec off s-1. The light

source used was LED of 0.09-0.1 W (ALAS, China), red 620-625 nm at 0.09 w per bulb and

blue 425-430 nm 0.1 W per bulb. It was used 400 red LEDs and 400 blue LEDs with a total

power consumption of 36 W in red and 40 W in blue with a combine total power of 76 W of

light in 56 L of algae suspension, and the power ratio was 1.35 W L-1. The LEDs were placed in

four square towers submerged in the alga suspension located at 8 cm from each other, a

transparent plexi glass protected the LEDs from water damage. The light emission per LED are

2143 µmol m-2 s-1 for blue and 1345 µmol m-2 s-1 for red. Light intensity was measure using a

quantum meter (Model MQ output in µmol m-2 s-1, Apogee Instruments Inc., Logan, UT, USA).

For the continuous light experiments we used the same type and brand of LEDs, 26 red and 26

blue with a light power of 2.34 W for red and 2.6 W for blue for a total of 4.94 W in 3.6 L

reactor. The power ratio was 1.37 W L-1. The LEDs were placed in the center of each container

with in a transparent plexi glass protection against water damage.

7.3.4 LIGHT DISPERSION

The LED technology is very efficien at producing light emissions in the photosynthetic spectra

from 1345 to 2143 µmol m-2 s-1 from a bulb of 0.09-0.1 W. This amount of light energy is more

than the algae species can process 50-250 µmol m-2 s-1, producing photo-inhibition and photo-

damage which ends in the reduction of the photosynthetic activity (Kim et al., 2006; Degen et al.,

2001; Ogbonna et al., 1999; Merchuk et al., 1998; Fernandez-Sevilla et al., 1998; Molina –Grima

et al., 1996, Garcia-Sanchez et al., 1996; Qiang and Richmond, 1994; Aiba, 1982; Bannister,

1979; Eppley and Coatsworth, 1966). The dispersion of light occurs in a short distance, the

distance from the bulb to the algal in suspension is 7 mm, sufficient distance to reduce the light

intensively from 2143 µmol m-2 s-1 to a power lower than 300 µmol m-2 s-1 which is where the

Page 143: Effects of Light, Co2 and Reactor Design on Growth of Algae

129 

 

photo inhibition process begins. Thus the resulting light intensity is optimal for alga phototropic

growth under artificial conditions (Ogbonna et al., 1999).

7.3.5 MIXING, QUANTIFYING CARBON SOURCE UNDER AUTOTROPHIC GROWTH

AND TEMPERATURE MEASUREMENTS

For carbon supply and gentle mixing we used in total three turbines (Model AC-9904, Resun,

China); For the flashing experiments we used one AC-9904 for each reactor, the air volume we

introduced was 9 L min-1 in 56 L, the air ratio was 161 mL of air L-1 of algae suspension min-1.

For the continuous light experiments we used one AC-9904 with volume of 9 L min-1, however

we only used two of the four outputs for the algae in this experiments, in total we used 4.5 L min-

1 in 7.2 L of algae suspension, the air ratio was 625 mL of air L-1 min-1 in the growing media.

The CO2 concentration in the two experiments ranged from 0.0323% to 0.0428% and growing

media temperature was 23°C + 2 measured by mini Infrared thermometer (Model Fluke 62,

Fluke Corporation, Everett, Washington, USA).

7.4.5 CULTURE MEDIUM

Comercial fertilizer (Solucat 25-5-5, Atlantica Agricola, Villena, Spain) was dissolved at 0.8 g in

1 L of distilled water. The fertilizer has the following composition; nitrogen 25% in ammonia

14,2% plus nitrates 10.8%. P2O5 5%, K2O 5%, Fe 0.020%, Mn 0.010%, Zn 0.002%, B 0.010%,

Cu 0.002%. Electrical conductivity was monitoring at 0.95 ms +- 0.01 mS using a EC sensor

(Model HI 991301, Hanna, Michigan, USA).

7.4.6 BIOMASS DETERMINATION

For the flashing light experiments, we used the biomass colorimetric determination methodology

to calculate the life standing biomass of the treatments. We used the spectrophotometer

(Shimadzu UV-1600, Shimadzu Scientific Instruments Inc, Columbia, USA) to prepare a

calibration curve, and measured the absorbance at 680 nm. A linear regression was made

Page 144: Effects of Light, Co2 and Reactor Design on Growth of Algae

130 

 

resulting in the following equation for Nannochloris sp Y= 0.8586x + 0.1464 with a R2 = 0.9854

and for Scenedesmus sp Y= 0.81x + 0.0461 with a R2 = 0.9813.

For the continuous light experiments we made a regression between biomass and cell density. To

measure the biomass we used the dry weight methodology that consist; dry a 15 mL glass tubing

until reach constant weight, fill the tubing with 10 mL of the algae media from the experiment,

centrifuge the tubing at 4000 rpm for 15 min, separate the water residues in the tubing, dry the

tubing with the algae in an oven at 105 oC until reaches constant weight, weight the tubing with

the algae substrate and weight the tubing without the algae, the resulting value is the algae

biomass in 10 mL. We repeated this procedure in duplicate each time we measured algae

biomass. For the cell density we used a microscope with a Neubauer chamber and we followed

the procedure for cell counting. A regression between cell density and biomass resulted in the

following equation for Nannochloris sp Y= 9*10-14 X2 + 2*10-05 X – 3.0797 with R2 =

0.9997and for Scenedesmus sp Y= 8*10-05 X – 117.94 with R2 = 1. The equations were used to

determine standing biomass from cell density.

7.4.7 EXPERIMENTAL DESIGN AND STATISTICAL ANALYSIS

The experiment measured standing biomass for two algae; one marine Nannochloris sp and one

fresh water Scenedesmus sp growing under daily solar radiation and 12 h artificial light; red and

blue LEDs flashing at 10 HZ, and growing under daily solar radiation and 12 h continuous

lighting red and blue LEDs. The algae were grown for 13 d in both experiments in a mediun of

Solucat fertilizer 0.8 g L-1, for the flashing experiment the light power ratio was at 1.35 W L-1,

the air intake 161 mL L-1 and for the continuous light experiment the light power ratio was 1.37

W L-1 and the air intake 625 mL L-1. The results were analyzed using statistical methods such

standard deviation, coefficient of variance, and one way ANOVA and T test using Biostatistics

1.0.

Page 145: Effects of Light, Co2 and Reactor Design on Growth of Algae

131 

 

7.5 RESULTS AND DISCUSSION

7.5.1 STANDING BIOMASS UNDER DIFFERENT LIGHT

The marine algal Nannochloris sp reached a standing biomass of 1041 mg L-1 under flashing

treatment, this overall biomass production is higher than the standing biomass reached under

continuous light treatment 892 mg L-1. The fresh water algal Scenedesmus sp reached a standing

biomass of 1130 mg L-1 under flashing light, the same algal under continuous light reached a

lower standing biomass 1018 mg L-1. In both algae species the flashing light treatments

developed more biomass production during the 13 d of the study than the continuous light. See

figure 7.1 for details of standing biomass during production during the experiment.

Figure 7.1. Standing biomass production under two light treatments for a marine algal Nannochloris sp and a fresh water alga Scenedesmus sp.

Page 146: Effects of Light, Co2 and Reactor Design on Growth of Algae

132 

 

7.5.2 DAILY BIOMASS PRODUCTION

The daily biomass production have a high index of fluctuation because the treatments include 13

daily solar radiations, and depending of the weather conditions the reactor receives more or less

light during the day photoperiod, and as a result in more or less photosynthesis ratio and biomass

production. The coefficient of variation for the daily biomass production under flashing light was

0.5 for Nannochloris sp and 0.8 for Scenedesmus sp, in continuous light the coefficient of

variation was 0.7 for Nannochloris sp and 1.0 for Scenedesmus sp.

There are statistical differences between flashing and continuous light treatment in the alga

Nanochloris sp resulting in ANOVA with F = 32.8889 and p = 0.0002, and T test with T = 8.967

and p = 0.000. Hovewer no statistical differences between treatments in Scenedesmus sp and

between algae species were found. Flashing and continuous light treatments induced different

biomass production for Nannochloris sp but not for Scenedesmus sp.

7.5.3 CHANGES IN POPULATION DENSITY

The marine alga Nannochloris sp developed the highest population density of this study under

continuous light 5.69*107 cells mL-1, under flashing light the cell density was 2.43*107 cell mL-1.

These results are contrary to these of the fresh water alga Scenedesmus sp which developed more

cell density under flashing light 1.65*107 cells mL-1 than under continuous light 1.42*107 cells

mL-1. See Figure 7.2 for details.

Page 147: Effects of Light, Co2 and Reactor Design on Growth of Algae

133 

 

Figure 7.2. Population density of two algae under two light treatments.

The growth rate of Nannochloris sp under flashing light treatments was 2.08 in 13 d, and the

average daily growth rate 0.160 and under continuous light the growth rate was 2.03 in 13 d, and

the average daily growth rate 0.196. There is statistical difference with F = 6.2567 and P =

0.0314 during the first 6 days of incubation in the growth rate of Nannochloris sp under flashing

light and continuous light. In the other hand Scenedesmus sp growth rate under flashing light was

1.48 and the daily average 0.148, and under continuous light the growth rate was 2.04 and the

daily average 0.194. No statistical differences was found in the growth rate of Scenedesmus sp

under flash light and continuous light. The growth rate of both algae fluctuate from 1.48 to 2.08

during the 13 d of incubation. In all two treatments Nannochloris sp grew faster and developed

more cell density than Scenedesmus sp because the cell size of Nannochloris sp is 2.5-3.2 µm in

diameter, smaller than the cell size of Scenedesmus sp 7 µm in width and 16 µm in lenght. See

Figure 7.3 for analysis of cell size.

Page 148: Effects of Light, Co2 and Reactor Design on Growth of Algae

134 

 

Figure 7.3. Photograph analysis of Nannochloris sp and Scenedesmus sp include size and some organelles.

The cell morphology of Nannochloris sp is round 2.5-3.2 µm in diameter, and the nucleus is 1.5

µm in diameter. The Scenedesmus sp cell morphology is capsule, the alga tipycally are grouped

in 2-4 cells, the cell size is 16 µm in length, 7 µm in width and the nucleus is 4 µm in diameter.

The cell has a recognisable inner nucleus of 2 µm in diameter, chloroplast structure and flagella

can be recognized as well in some cells, the nucleus of Scenedesmus sp is bigger than

Nannochloris sp cells.

7.5.4 CELL WEIGHT

During the 13 d experiments the cell weight was calculated using the cell density measurements

and the cell biomass, a total of 8 samples were taken for cell density and 8 samples for cell

biomass in each treatment. Table 7.1 shows the cell weight for the two algae species under two

light treatments.

Page 149: Effects of Light, Co2 and Reactor Design on Growth of Algae

135 

 

Table 7.1. Cell weight of two algae species under two light treatments.

Cell weight (g) for two algae under two treatments 

Repetition Nannochloris sp  Scenedesmus sp 

Flashing  Continous  Flashing  Continous 

1  5.33E‐11  2.84E‐11  1.15E‐10  1.62E‐11 

2  6.10E‐11  2.62E‐11  1.09E‐10  3.36E‐11 

3  5.49E‐11  2.50E‐11  7.41E‐11  3.93E‐11 

4  4.56E‐11  2.37E‐11  8.50E‐11  4.81E‐11 

5  4.43E‐11  2.12E‐11  8.12E‐11  6.05E‐11 

6  4.39E‐11  1.95E‐11  7.90E‐11  6.35E‐11 

7  5.42E‐11  1.91E‐11  8.62E‐11  7.13E‐11 

8  4.13E‐11  1.76E‐11  5.08E‐11  7.17E‐11 

Mean  4.98E‐11  2.26E‐11  8.50E‐11  5.05E‐11 

Standar Deviation  6.94E‐12  3.82E‐12  2.00E‐11  1.98E‐11 

Coeficient of Variance  0.14  0.17  0.24  0.39 

The average cell weight of Scenedesmus sp under the flashing light is 8.50*10-11 g and for

continous light 5.05*10-11 g, almost double the weight of Nannochloris sp cell which under

flashing treatment reached 4.98*10-11 g and for continuous light 2.26*10-11 g. There is a

statistical difference in cell weight in Nannochloris sp between flashing and continuous light

with F = 94.2495, T = 9.708 and p = 0.000. Similar results in cell weight were obtained with

Scenedesmus sp with F = 11.9953, T = 3.463 and p = 0.004. The cell weight is higher under the

flashing conditions, not due to the light treatment, but could be as a result of the turbulences. The

flashing light reactor is a bigger reactor 56 L with an air ratio of 161 mL of air L-1, on the other

hand the continuous light reactor is 3.6 L with an air ratio of 625 mL of air L-1, this air ratio

produce more turbulence in the smaller reactor which affect directly the size of the cell.

Turbulences system develop smaller cell because of the increasing probability of mechanical

damage as the cell grow, therefore the differences of cell size in not produced by the light

treatment but the difference of turbulences of the two systems.

Page 150: Effects of Light, Co2 and Reactor Design on Growth of Algae

136 

 

7.6 DISCUSSION AND CONCLUSIONS

The standing biomass obtained in this study under flashing and continuous light after 13 days of

incubation for both algae Nannochloris sp and Scenedesmus sp were in the range of 892-1130

mg L-1, these biomass production is similar to those obtained by Bhola et al., 2010 and Liang et

al., 2009 which confirm that 12-15 days of incubation of Chlorella vulgaris produces 250 to

1700 mg L-1 of biomass. However there is a big potential in biomass production using flashing

light effect, a study made with C. vulgaris in a Taylor vortex reactor has reached a standing

biomass of 10000 mg L-1 in just 4 days of incubation (Kong et al., 2013). In relation with the size

of the algae and the population density; Nannochloris sp produces more cell density 2.43*107 to

5.69*107 cell mL-1 with a smaller cell size 2.5-3.2 µm in diameter and Scenedesmus sp produce

less cell density 1.42*107 to 1.65*107 cells mL-1 with a bigger cell size 7 µm width and 16 µm

length.

The growth rate of the two algae is different because they produce similar standing biomass

(14% differences) with different cell size (3x-7x); therefore the algal that has a smaller cell

(Nannochloris sp) produce more cell density in order to produce similar biomass. Thus

Nannochloris sp cell weight range from 2.26*10-11 to 4.98*10-11 g Cell-1 smaller weight than

Scenedesmus sp cell weight 5.05*10-11 to 8.50*10-11 g Cell-1.

The ANOVA test confirm that growing Nannochloris sp under flashing light produce more

biomass and growth rate during the first 6 days of incubation than continuous light. The daily

biomass production and growth rate in the alga Scendesmus sp for two different treatments

(continuous light and flashing light) resulted in no statistical difference. However we cannot

conclude that flashing light does not have any effect in more or less biomass production because

Scenedesmus sp produced more standing biomass under flashing light than under continuous

light treatment.

Page 151: Effects of Light, Co2 and Reactor Design on Growth of Algae

137 

 

This study found statistical difference in growth rate and daily biomass production for the marine

algae Nannochloris sp, however did not find any significant difference in daily biomass and

growth rate for the fresh water alga Scenedesmus sp. The two algae are morphologically

different; Nannochloris sp is smaller, produce more cell density and has a higher growth rate.

However the standing biomass of Scenedesmus sp under flashing light is 8.5% higher than the

standing biomass of Nannochloris sp, and for the continuous light is 14% higher. Both algae

increased their cell weight as the turbulence of the media decrease.

Page 152: Effects of Light, Co2 and Reactor Design on Growth of Algae

138 

 

7.8 REFERENCES

1. Aiba, S. 1982. Growth kinetics of photosynthetic microorganisms. Adv. Biochem. Eng. J. 23: 85-156.

2. Bannister, T.T. 1979. Quantitative description of steady state, nutrient-saturated algal growth, including adaptation. Limnol. Ocean. 24: 76-96.

3. Bhola, V., Desikan, R., Santosh, S. K., Subburamu, K., Sanniyasi, E., Bux, F. 2011. Effects of parameters affecting biomass yield and thermal behaviour of Chlorella vulgaris. J. Bioscien. Bioeng. 111-3: 377-382.

4. Degen, J., Uebele, A., Retze, A., Schmid-Staiger, U., Trosch, W. 2001. A novel airlift photobioreactor with baffles for improved light utilization through the flashing light effect. J. Biotechnol. 92: 89-94.

5. Eppley, R.W., Coatsworth, J.L. 1996. Culture of the marine phytoplankter. Dunaliella tertiolecta, with light-dark cycles. Arch. Microbiol. 55: 66-80.

6. Fathurrahman, L., Siti-Hajar, A.H., Wan-Nur-Sakinah, D., Nurhazwani, Z., Ahmad, J. 2013. Flashing ligh as growth stimulant in cultivation of green microalgae Chlorella sp utilizing airlift photobioreactor. Pak. J. Biol. Sci. 16-22: 1517-1523.

7. Fernandez-Sevilla, J.M., Molina-Grima, E., Garcia-Camacho, F., Acien-Fernandez, F.G., Sanchez-Perez, J.A. 1998. Photolimitation and photoinhibition as factors determining optimal dilution rate to produce eicosapentaenoic acid from cultures of the microalga Isochrysis galbana. Appl. Microbiol. Biotechnol. 50: 199-205.

8. Garcia-Sanchez, J.L., Sanchez-Perez, J.A., Garcia-Camacho, F., Fernandez-Sevilla, J.M., Molina-Grima, E. 1996. Optimization of light and temperature for growing Chlorella sp using response surface methodology. Biotechnol. Techn. 10: 329-334.

Page 153: Effects of Light, Co2 and Reactor Design on Growth of Algae

139 

 

9. Grobbelaar, J. U. 1991. The influence of light/dark cycles in mixed algal cultures on their productivity. Bioresour. Technol. 38: 189-194.

10. Janssen, M., Kuijpers, T. C., Veldhoen, B., Ternbach, M. B., Tramper, J., Mur, L. R. 1999. Specific growth rate of Chlamydomonas reinhardtii and Chlorella sorokiniana under medium duration light/dark cycles 13-87 s. J. Biotechnol. 70: 323-333.

11. Kim, J.P., Kang, C.D., Park, T.H., Kim, M.S., Sim, S.J. 2006. Enhanced hydrogen production by controlling light intensity in sulphur-deprived Chlamydomonas reinhardtii culture. Int. J. Hydrogen Energy. 31: 1585-1590.

12. Kong, B., Shanks, J.V., Vigil, R.D. 2013. Enhanced algal growth rate in a Taylor vortex reactor. Biotechnol. Bioeng. 110-8: 2140-2148.

13. Laws, E. A., Terry, K.L., Wickman, J., Chalup, M.S. 1983. A simple algal production system designed to utilize the flashing light effect. Biotechnol. Bioeng. 10: 2319-2335.

14. Lee, Y.-K. & Pirt, S. J. 1981. Energetics of photosynthetic algal growth: influence of intermittent illumination in short (40 s) cycles. J. Gen. Microbiol. 124: 43-52.

15. Liang, Y., Sarkany, N., Cui, Y. 2009. Biomass and lipid productivities of Chlorella vulgaris under autotrophic, heterotrophic and mixotrophic growth conditions. Biotechnol. Lett. 31: 1043-1049.

16. Merchuk, J. C., Ronen, M., Giris, S., Arad, S. M. 1998. Light/dark cycles in the growth of the red microalga Porphyridium sp. Biotechnol. Bioeng. 59-6: 705-713.

17. Molina-Grima, E., Fernandez-Sevilla, J. M., Sanchez-Perez, J. A., Garcia-Camacho, F. 1996. A study on simultaneous photolimitation and photoinhibition in dense microalgal cultures takin into account incident and averaged irradiances. J. Biotechnol. 45: 59-69.

Page 154: Effects of Light, Co2 and Reactor Design on Growth of Algae

140 

 

18. Myers, J., Graham, J. R. 1971. The photosynthetic unit in in Chlorella measured by repetitive short flashes. Plant Physiol. 48: 282-286.

19. Ogbonna, J.C., Soejima, T., Tanaka, H. 1999. An integrated solar and artificial light system for internal illumination of photobioreactors. J. Biotechnol. 70: 289-297.

20. Park, K. H., Lee, C. G. 2001. Effectiveness of flashing light for increasing photosynthetic efficiency of microalgal cultures over a critical cell density. Biotechnol. Bioprocess Eng. 6: 189-193.

21. Park. K.H., Lee, C.G. 2000. Optimization of algal photobioreactors using flashing lights. Biotechnol. Bioprocess Eng. 5: 186-190.

22. Philiph, J.N., Myers, J. 1954. Growth rate of Chlorella in flashing light. Plant Physiol. 29: 152-161.

23. Pirt, S. J. 1986. The thermodynamic efficiency (quantum demand) and dynamics of photosynthetic growth. New Phytologist. 102-1: 3-37.

24. Pirt, S.J., Lee, Y.K., Richmond, A., Pirt, M.W. 1980. The photosynthetic efficiency of Chlorella biomass growth with reference to solar energy utilisation. J. Chem. Tech. Biotechnol. 30: 25-34.

25. Qiang, H., Richmond, A. 1994. Optimizing the population density in Isochrysis galbana grown outdoors in a glass column photobioreactor. J. Appl. Phycol. 6: 391-396.

26. Walsh, P., Legendre, L. 1988. Photosynthetic responses of the diatom Phaeodactylum tricornutum to high frequency light fluctuations simulating those induced by sea surface waves. J. Plankton Res. 10-5: 1077-1082.

Page 155: Effects of Light, Co2 and Reactor Design on Growth of Algae

141 

 

27. Vejrazka, C., Janssen, M., Streefland, M., Wijffels, R. H. 2012. Photosynyhetic efficiency of Chlamydomonas reinhardtii in attenuated flashing light. Biotechnol. Bioeng. 109: 2567-2574.

28. Yago, T., Arakawa, H., Fukui, K., Okubo, B., Akima, K., Takeichi, S., Okumura, Y., Morinaga, T. 2012. Effects of flashing light from light emitting diodes (LEDs) on growth of the microalga Isochrysis galbana. African J. Microbiol. Research. 6-30: 5896-5899.

Page 156: Effects of Light, Co2 and Reactor Design on Growth of Algae

142 

 

 

CHAPTER 8 EFFECT OF CARBON DIOXIDE CONCENTRATION ON THE

GROWTH RESPONSE OF CHLORELLA VULGARIS UNDER FOUR DIFFERENT LED ILUMINATION

Sebastián Mejía Rendón Advisor: R. Paul Voroney1

International Journal of Biotechnology for Wellness Industries, 2013, 2, 125-131

8.1 ABSTRACT

This experiment examined the growth response of Chlorella vulgaris exposed to CO2

concentrations increasing from ambient to 8-9% and under white, blue, red and red+blue lights

after 15 days incubation. Biomass production increased with increasing CO2 concentrations

under all light sources. The highest biomass production, 1.59 g L-1, was obtained when the algae

were supplied with 8-9% CO2 and exposed to white light. Biomass production under blue, red

and red+blue light was 1.53 g L-1, 0.45 g L-1 and 1.27 g L-1, respectively. The research suggests

that C. vulgaris is not able to adapt production of its photosynthetic pigments to absorb light

sources different that it is normally has evolved to.

Keywords: Chlorella vulgaris, Photobioreactor, Biomass production, CO2 concentration,

Artificial light.

8.2 INTRODUCTION

Phototrophic algal growth requires light, mineral nutrients, water and an inorganic source of

carbon (CO2). While there are algae, such as Chlorella sp, that can grow relatively rapidly under

the ambient air concentration of CO2 (0.037%) (Usui and Ikenouchi, 1997; Hirata et al., 1996),

maximizing biomass production for the purpose of CO2 capture requires that higher

concentrations of CO2 be provided. Algae have been grown in closed atmospheres at high CO2

Page 157: Effects of Light, Co2 and Reactor Design on Growth of Algae

143 

 

concentrations (10-20% and higher) with the objective of CO2 fixation (Cheng et al., 2006;

Brown, 1996; Zeiler et al., 1995). The alga C. vulgaris is a good candidate for biomass

production under high CO2 concentrations because it is able to fix up to 74% of the original CO2

with only 2 seconds of CO2 residence time (Keffer and Kleinheinz, 2002).

Recently designs of closed photo-bioreactors for the purpose of enhancing light-use efficiency,

CO2 fixation, as well as biomass production, have received more attention because algae are

efficient photosynthetic organisms (Kajiwara et al., 1997) and have the potential to reduce

atmospheric CO2 levels (Yoshihara et al., 1996), or reduce emissions from a gas or coal power

plant. Thus C. vulgaris produce compounds of economical value such as antioxidant (β-

carotenes), natural colorants, oils such as omegas 3, 6 and 9, and proteins and carbohydrates.

This study evaluated standing biomass production of the alga C. vulgaris growing under normal

and elevated CO2 concentrations and exposed to 4 different light sources.

8.3 MATERIALS AND METHODS

8.3.1 MICROORGANISM

The algal specie, Chlorella vulgaris (UTEX 26), was obtained from the culture collection of

algae at the University of Texas at Austin (UTEX, 205 W, 24th St., Austin, Texas, TX 78712,

USA). The culture was established in 1 L batch for 15 d and then transferred to 4 L reactors. The

population density was established using direct microscopic counting techniques. The population

density at the start of the experiment was set at 4.5- 5*105 cells mL-1 by adjusting the volume of

medium in the reactors.

8.3.2 LIGHT SOURCE

Sources providing white, blue, red, and red-blue light were evaluated. They were 0.1 W LEDs

obtained from ALAS™ (China) with the following emission parameters: red, 620-625 nm

Page 158: Effects of Light, Co2 and Reactor Design on Growth of Algae

144 

 

wavelength at 1345 µmol m-2 s-1 intensity; blue, 425-430 nm wavelength at 2143 µmol m-2 s-1;

white, 380-760 nm wavelength at 1838 µmol m-2 s-1 intensity.

Each lighting system contained 52 LEDs fixed into a clear square plexiglass tube (dimension 4*4

cm and 33 cm in length) and placed in the centre of the 4-L culture flask (internal diameter 12

cm). The illumination system was powered by a transformer (Model SA 201-3485, ASTEC,

California, USA) supplying 12 V and 8 Amp. To ensure that external illumination would not

affect the experiment, the laboratory was kept dark (0 µmol m-2 s-1) throughout the incubation

period.

Light intensity was monitored during the incubation using 2 different light meters:

photosynthetic light was measured using a quantum meter (Model MQ output in µmol m-2 s-1,

Apogee Instruments Inc., Logan, UT, USA) and visible light was measured using a lux meter

(Model MS6610 output in Lux, VIA Instruments, Shanghai, China)

8.3.3 LIGHT DISPERSION

The LEDs light emissions are very powerful; a bulb of 0.1 W produces light emissions in the

photosynthetic spectra from 1300 to 2140 µmol m-2 s-1 depending in the wavelength. This

amount of energy surpasses the maximum photosynthetic light absorption from algae species that

is 50-250 µmol m-2 s-1 causing photo-inhibition and photo-damage (Kim et al., 2006; Degen et

al.,2001; Fernandez-Sevilla et al., 1998;. Merchuk et al., 1998) Nevertheless the dispersion of

this light happens in a short distance, thus the algal in the present study receives lower light

emissions than the maximum value for photosynthetic growth. The distance from the bulb to the

algal in suspension is 7 mm enough distance to reduce the light intensively lower than 300 µmol

m-2 s-1 that is where the photo inhibition process begins, thus the resulting light intensity is

perfect for algal phototropic growth under artificial conditions (Ogbonna et al., 1999). See Table

8.1 for more details about light dispersion versus distance in water measured during the course of

this experiment with the Apogee quantum meter.

Page 159: Effects of Light, Co2 and Reactor Design on Growth of Algae

145 

 

Table 8.1. LED light dispersion and light reducing intensity through distance.

Light dispersion from the source

Light emission in (µmol m-2 s-1)

Distance from source (mm)

Blue Red White

0 2143 1345 1828 1 568 225 413 5 206 173 186 10 95 72 103

8.3.4 PHOTOBIOREACTOR AND CO2 SUPPLY

The photobioreator is a bubble-column glass container of 36 cm length, 12 cm diameter, and

volume 4 L. Carbon dioxide at concentrations of 0.035% (350 ppm ± 50), 1.1%, 3.7% and 8.5%

± 0.18% was supplied at a rate of 864 mL min-1 L-1 of algal suspension from a compressed air

tank equipped with a mass flow regulator (Model 191 AR-60, Gentec Corporation, Shanghai,

China). Two turbines, (Model ACO-9720, Hailea, Hailea Industrial Zone, Guangdong, China)

each with an output of 30 L min-1, were used to mix the algal suspension. The concentration of

dissolved CO2 in the inlet gases was measured at 5 s intervals with a CO2 monitor (Model 7001,

Telaire-General Electric, California, USA). Room temperature was measured with a temperature

monitor (Model 7001, Telaire-General Electric, California, USA). Periodically during the

experiment the algal suspension temperature was measured with an infrared thermometer (Model

Fluke 62, Fluke Corporation, Everett, Washington, USA), Both room temperature and algae

suspension temperature were about 22°C + 0.9.

8.3.5 CULTURE MEDIUM

The culture medium was prepared using a commercial synthetic fertilizer (Solucat 25-5-5,

Atlantica Agricola, Villena, Spain) by dissolving 0.8 g in 1 L distilled water. At the start of each

Page 160: Effects of Light, Co2 and Reactor Design on Growth of Algae

146 

 

experiment the culture medium contained 114 mg L-1 ammonium-N, 86.4 mg L-1 nitrate-N, 40

mg L-1 phosphate, 40 mg L-1 potash, 160 μg L-1 iron, 80 μg L-1 manganese, 80 μg L-1 boron, 16

μg L-1 zinc, and 16 μg L-1 copper; the electrical conductivity was 0.95 ± 0.01 mS measured using

a EC sensor (Model HI 991301, Hanna, Michigan, USA).

8.3.6 EXPERIMENTAL DESIGN AND STATISTICAL ANALYSIS

Each experiment was set out in 24 units of 4 L photobioreactors, exposed to four different light

treatments (white, blue, red and red-blue) and 6 replicates, and incubated for 15 d. Each

photobioreactor contained 3 L of culture medium. The experiment was repeated for the four

different CO2 concentrations.

8.3.7 BIOMASS ANALYSIS

Colorimetric determination methodology was used to calculate the standing biomass production

at the end of the incubation. Light absorption of each sample was measured at 680 nm with a

spectrophotometer (Shimadzu UV-1600, Shimadzu Scientific Instruments Inc, Columbia, USA).

For the calibration of this methodology seven samples with different biomass concentrations of

C. vulgaris, ranging from 0.25 to 3.49 g L-1 were used to establish a relationship between

standing biomass an absorbance. The resulted equation Y = 0.7015x + 0.0362 with R2 of 0.9987

was used to determine the biomass from 96 replicates in this study.

8.3.8. STATISTICAL ANALYSIS

The results were analyzed using ANOVA test with the software Biostatistics 1.0, and Stats Pad

version 1.3 at α≤ 0.01. An ANOVA was performed for different CO2 concentrations as well as

for different light treatments.

Page 161: Effects of Light, Co2 and Reactor Design on Growth of Algae

147 

 

8.4 RESULTS AND DISCUSSION

8.4.1 STANDING BIOMASS PRODUCTION UNDER INCREASING CO2

CONCENTRATIONS

Standing biomass of C. vulgaris increased with increasing concentrations of CO2 under the four

different light sources (Figure 8.1). The highest biomass production, 1.59 g L-1, was found when

the algal culture were supplied with 8.5% CO2 and exposed to white light. Biomass production

under blue, red and red+blue light was 1.53 g L-1, 0.45 g L-1 and 1.27 g L-1, respectively. An

experiment growing Chlorella sp under increasing concentrations of CO2 found that the standing

biomass increased from 0.5 to 5.7 g L-1 reaching the highest standing biomass when the CO2

reached 10% (Sung et al., 1998). A similar study growing Chlorella sp at different CO2

concentrations resulted in standing biomass of 2 g L-1 at 10% CO2 (Maeda et al.,1995). Other

study with Chlorella showed a standing biomass of 3 g L-1 when the algal was grown at 10%

CO2, also good growth was reported with Chlorella sp at CO2 concentrations from 10 to 50%

(Sung et al., 1999), reaching 2 g L-1 when the CO2 concentration range from 5 to 40% (Sakai et

al., 1995). Standing biomass of 2 g L-1 also was obtained growing Chlorella sp at 5% CO2 (Ryu,

et al., 2009), concentration of 2% and 10% CO2 has resulted in biomass synthesis of 1.67 to 1.5 g

L-1 after 6 days of incubation (Chiu et al., 2011). Several studies conducted with C. vulgaris have

reported a typical standing biomass between 0.25 g L-1 and 1.7 g L-1 under phototropic growth

after 12-15 days of incubation (Bhola et al., 2010; Liang et al., 2009), growing Chlorella sp

under digested manure had produced standing biomass of 1.7 g L-1 after 21 days of incubation

(Pitman et al., 2011; Wang et al., 2010). The lowest levels of standing biomass 0.21 g L-1 were

obtained growing C. vulgaris in waste water (Chinnasamy et al., 2009), however using artificial

waste water C. vulgaris had developed standing biomass of 1.6 g L-1 after 11 days of incubation

(Feng and Zhang, 2011).

Page 162: Effects of Light, Co2 and Reactor Design on Growth of Algae

148 

 

0

0.2

0.4

0.6

0.8

1

1.2

1.4

1.6

1.8

0.035% 1.10% 3.70% 8.5%

Stan

ding Biomass (g /L)

Changes in Standing Biomass Under Different Light and CO2 Concentrations 

Red

Blue

Red‐Blue

White

Figure 8.1. Changes in standing biomass under different concentration of CO2 and different wavelength illumination.

8.4.2 STATISTICAL ANALYSIS OF THE STANDING BIOMASS

The present statistical analysis is a two way ANOVA comparing two variables; the four CO2

concentration treatments and the four light treatments. The ANOVA test results in a F in the

samples = 13319, columns = 38194 and interaction = 2480 with p = 0.0000, these results show

significant differences in the two variables CO2 and light. In addition, a one way ANOVA was

used to compare individual treatments of CO2 and light to one another. For example; the algae

growing in four CO2 concentration treatments under white light to analysis possible differences

in the standing biomass as the CO2 concentration changes. The second analysis is comparing the

same CO2 concentration under different light treatment with the objective of determining if the

wavelength of the light affects standing biomass. Table 8.2 compares the effect of CO2

concentration in the standing biomass in four different type of illumination.

Page 163: Effects of Light, Co2 and Reactor Design on Growth of Algae

149 

 

Table 8.2. Analysis of variance of the standing biomass for four different CO2 concentration under costant light sources.

ANOVA different CO2 concentration at p = 0.01

Color Red Blue Red-Blue White F table

Different CO2 3966 12466 13256 11174 4.94

The ANOVA shows a statistical difference at p = 0.01 in the standing biomass of the algae C

vulgaris growing under increasing CO2 concentrations (0.035%, 1.1%, 3.7% and 8.5% CO2) and

4 different light wavelengths. The F value from the experiment are between 1117 and 13256 and

the F of the table at p = 0.01 is 4.94. The statistical analysis confirms that the differences in

standing biomass production with increasing CO2 concentrations and under the four different

light sources are significant. For the next ANOVA test (Table 8.3) we leave the CO2

concentration constant to see changes in the standing biomass if the light wavelength changes

and the light power remains constant.

Table 8.3. Analyses of the standing biomass if the light regime changes and the CO2 concentration remains constant.

ANOVA different wave length illumination at p = 0.01

CO2 Concentration 0.0350% 1.10% 3.70% 8.5% F table

Different color 1579 3983 2308 7405 4.94

The ANOVA for light source shows statistical differences at p = 0.01. The tests for the four CO2

concentrations gave an F ranging from 740 to 3982 and the F of tables is 4.94. There is statistical

difference in C. vulgaris standing biomass when is grown under different concentration of CO2

and exposed to light sources of different wavelength.

Page 164: Effects of Light, Co2 and Reactor Design on Growth of Algae

150 

 

8.4.3 LIGHT SOURCE EFFECTS ON BIOMASS PRODUCTION

Exposure of the algae to white LED light and supplied with 8.5% CO2 concentration resulted in

the highest standing biomass of this study 1.6 g L-1, even higher than 0.16 g L-1 obtained in a

study growing Scenedesmus dimorphus at 10% CO2 (Lunka and Bayless, 2013). Different results

were obtained growing Nannochloropsis sp under LED illumination, the results of this study

show that blue 470 nm developed the highest growth rate follow by white, green 550 nm and red

680 nm (Das, et al., 2011). Contrary to the rest of the lights treatments, biomass production

under red light was highest 0.57 g L-1 when the cultures were exposed to 3.7% CO2 and not at

the highest CO2 concentration 8.5% where biomass was 0.44 g L-1. The red 625 nm always

developed less biomass than the rest of the lights treatments, these results confirm the study

made with Isochrysis galbana which demonstrated that a shorter wavelength such as blue 460

nm is more photosynthetic efficient than a longer wavelength as red 670 nm (Jeon, et al., 2013).

However a study made with C. vulgaris growing in synthetic waste water and under different

LED illumination, conclude that red 660 nm developed higher biomass 0.28 g L-1 than white

0.25 g L-1, yellow 590 nm 0.21 g L-1, purple 410 nm 0.16 g L-1 blue 460 nm 0.15 g L-1 and green

550 nm 0.1 g L-1 (Yan, et al., 2013). Red 660-680 nm match better the C. vulgaris absorption

peak than red 625 nm, however it was expected C. vulgaris adaptation to this light spectra by

modifying its chlorophylls or by producing more complementary pigments such as carotenoids in

the absorption peak of 620-630 nm, little evidence of light spectra adaptation is suggested from

the results of this study.

The highest biomass production in cultures exposed to blue light 1.5 g L-1 were obtained when

the algal grew at 8.5% of CO2. However biomass production under blue light was higher than the

other light sources tested in this study for CO2 concentrations of 1.1% and lower when the

standing biomass is less than 0.6 g L-1. Since the photosynthetic part of algae are the chloroplast,

light reaching the rest of the algae body is not been used for photosynthesis, therefore increases

in standing biomass increase light shadowing. The shadowing effect is light blocked and not

used for photosynthetic process by algae located near to the light source resulting in less light

Page 165: Effects of Light, Co2 and Reactor Design on Growth of Algae

151 

 

reaching those algae farther away from the light source. This effect cause loss of photons,

therefore the amount of energy lost is greater at lower wavelength when the light photon has

more energy (Das et al., 2011). Since blue 425 nm light has more energy per photon (6.68*10-19

J) than do the other light sources for example red 625 nm (3.18*10-19 J), losses of light energy

caused by light shadowing would be greatest under blue light exposure. From the data in Table

8.4, the limit where the biomass production responds less to blue light switches between 0.6 to

0.8 g L-1. At this algal biomass concentration, the light source combination of red-blue

performed better than blue light alone because the red light losses for shadowing effect are less

than blue, and red light complements the blue light resulting in a higher standing biomass. White

light travels further and would have reached more distant algae, therefore would perform better

when the standing biomass is greater than 0.8 g L-1.

Table 8.4. Mean standing biomass production under increasing CO2 concentrations and exposed to four different light sources.

Standing Biomass (g L-1)

CO2 /Light source 0.035% 1.10% 3.70% 8.5%

Red 0.148 + 0.009 0.254 + 0.008 0.571 + 0.006 0.446 + 0.005

Blue 0.429 + 0.007 0.610 + 0.006 0.828 + 0.013 1.531 + 0.014

Red-blue 0.309 + 0.003 0.483 + 0.003 1.003 + 0.009 1.271 + 0.016

White 0.296 + 0.006 0.548 + 0.006 1.045 + 0.014 1.592 + 0.021

Note: + Standard deviation.

8.5. CONCLUSIONS

Production of C. vulgaris biomass increased when supplied with increasing CO2 concentrations

up to 8.5% under the four light sources. Growth of the algae was better under blue light when

algae were supplied with lower CO2 concentrations and the standing biomass was low. The

results of this study show that C. vulgaris does not adapt production of their photosynthetic

pigments to absorb light from a wavelength spectrum different from one that they would

normally be exposed to.

Page 166: Effects of Light, Co2 and Reactor Design on Growth of Algae

152 

 

Further studies with more CO2 concentration are recommended to establish the limit where CO2

is no longer a factor for increasing biomass production. It is further recommended to compare C.

vulgaris growth under red light at 660 nm, 670 nm with that under red at 625 nm.

ACKNOWLEDGEMENTS

This research was funded by the grants of Professor Paul Voroney at the School of Enviromental

Sciences, University of Guelph, Ontario Canada.

Page 167: Effects of Light, Co2 and Reactor Design on Growth of Algae

153 

 

8.6 REFERENCES

1. Bhola, V., Desikan, R., Santosh, S. K., Subburamu, K., Sanniyasi, E., Bux, F. 2011. Effects of parameters affecting biomass yield and thermal behaviour of Chlorella vulgaris. J. Biosci. Bioeng. 111-3: 377-382.

2. Brown, L. M. 1996. Uptake of carbon dioxide from flue gas by microalgae. Energy. Convers. Mgmt. 37-6-8: 1363-1367.

3. Cheng, L., Zhang, L., Chen, H., Gao, C. 2006. Carbon dioxide removal from air by microalgae cultured in a membrane-photobioreactor. Sep. Purif. Technol. 50: 324-329.

4. Chinnasamy, S., Ramakrishnan, B., Bhatnagar, A., Das, K. 2009. Biomass production potential of a wastewater alga Chlorella vulgaris ARC 1 under elevated levels of CO2 and temperature. Int. J. Mol. Sci. 10: 518-532.

5. Chiu, S.Y., Kao, C.Y., Huang, T.T., Lin, C.J., Ong, S.C., Chen, C.D., Chang, J.S., Lin, C.S. 2011. Microalgal biomass production and on-site bioremediation of carbon dioxide, nitrogen oxide and sulfur dioxide from flue gas using Chlorella sp cultures. Bioresour. Technol. 102: 9135-9142.

6. Das, P., Lei, W., Aziz, S.S., Obbard, J.P. 2011. Enhanced algae growth in both phototrophic and mixotrophic culture under blue light. Bioresour. Technol. 102: 3883-3887.

7. Degen, J., Uebele, A., Retze, A., Schmid-Staiger, U., Trosch, W. 2001. A novel airlift photobioreactor with baffles for improved light utilization through the flashing light effect. J. Biotechnol. 92: 89-94.

8. Feng, Y., Li, C., Zhang, D. 2011. Lipid production of Chlorella vulgaris cultured in artificial wastewater medium. Bioresour. Technol. 102: 101-105.

Page 168: Effects of Light, Co2 and Reactor Design on Growth of Algae

154 

 

9. Fernandez-Sevilla, J.M., Molina-Grima, E., Garcia-Camacho, F., Acien-Fernandez, F.G., Sanchez-Perez, J.A. 1998. Photolimitation and photoinhibition as factors determining optimal dilution rate to produce eicosapentaenoic acid from cultures of the microalga Isochrysis galbana. Appl. Microbiol. Biotechnol. 50: 199-205.

10. Hirata, S., Hayashitani, M., Taya, M., Tone, S. 1996. Carbon dioxide fixation in batch culture of Chlorella sp using a photobioreactor with a sunlight-collection device. J. Ferment. Bioeng. 81-5: 470-472.

11. Jeon, H., Lee, J., Cha, M. 2013. Energy efficient growth control of microalgae using photobiological methods. Renew. Energy. 54: 161-165.

12. Kim, J.P., Kang, C.D., Park, T.H., Kim, M.S., Sim, S.J. 2006. Enhanced hydrogen production by controlling light intensity in sulphur-deprived Chlamydomonas reinhardtii culture. Int. J. Hydrogen Energ. 31: 1585-1590.

13. Kajiwara, S., Yamada, H., Ohkuni, N., Ohtaguchi, K. 1997. Design of the bioreactor for carbon dioxide fixation by Synechococcus PCC7942. Energy Convers. Mgmt. 38: 529-532.

14. Keffer, J. E., Kleinheinz, G. T. 2002. Use of Chlorella vulgaris for CO2 mitigation in a photobioreactor. J. Ind. Microbiol. Biotechnol. 29: 275-280.

15. Liang, Y; Sarkany, N; Cui, Y. 2009. Biomass and lipid productivities of Chlorella vulgaris under autotrophic, heterotrophic and mixotrophic growth conditions. Biotechnol. Lett. 31: 1043-1049.

16. Lunka, A.A., Bayless, D.J. 2013. Effects of flashing light-emitting diodes on algal biomass productivity. J. Appl. Phycol. DOI 10.1007/s10811-013-0044-1

Page 169: Effects of Light, Co2 and Reactor Design on Growth of Algae

155 

 

17. Maeda, K., Owada, M., Kimura, N., Omata, K., Karube, I. 1995. CO2 fixation from the flue gas on coal-fired thermal power plant by microalgae. Energy Convers. Mgmt. 36-6-9: 717-720.

18. Merchuk, j.C., Ronen, M., Giris, S., Arad, S. 1998. Light/dark cycles in the red growth of the Microalga Porhyridium sp. Biotechnol. Bioeng. 59-6: 705-713.

19. Ogbonna, J.C., Soejima, T., Tanaka, H. 1999. An integrated solar and artificial light system for internal illumination of photobioreactors. J. Biotechnol. 70-1-3: 289-297.

20. Pitman, J.K., Dean, A.P., Osudenko, O. 2011. The potential of sustainable algal biofuel production using wastewater resources. Bioresour. Technol. 102: 17-25.

21. Ryu, H. J., Oh, K, K., Kim, Y, S. 2009. Optimization of the influential factors for the improvement of CO2 utilization efficiency and CO2 mass transfer rate. J. Ind. Eng. Chem. 15: 471-475.

22. Sakai, N., Sakamoto,Y., Kishimoto, N., Chihara, M., Karube, I. 1995. Chlorella strains from hot springs tolerant to high temperature and high CO2. Energy Convers. Mgmt. 36-6-9: 693-696.

23. Sung, K. D., Lee, J. S., Shin, C. S., Park, S. C. 1999. Isolation of a new highly CO2 tolerant fresh water microalga Chlorella sp KR-1. Renew. Energ. 16: 1019-1022.

24. Sung, K. D., Lee, J. S., Shin, C. S., Park, S. C., Choi, M. J. 1998. CO2 fixation by KR-1 and its cultural characteristics. Bioresour. Technol. 68: 269-273.

25. Usui, N., Ikenouchi, M. 1997. The biological CO2 fixation and utilization project by RITE(1) Highly-effective photobioreactor system. Energy Convers. Mgmt. 38: 487-492.

Page 170: Effects of Light, Co2 and Reactor Design on Growth of Algae

156 

 

26. Wang, L., Li, Y., Chen, P., Min, M., Chen, Y., Zhu, J., Rua, R.R. 2010. Anaerobic digested dairy manure as a nutrient supplement for cultivation of oil-rich green microalgae Chlorella sp. Bioresour. Technol. 101-8: 2623-2628.

27. Yan, C., Zhao, Y., Zheng, Z., Luo, X. 2013. Effects of various LED light wavelengths and light intensity supply strategies on synthetic high-strength wastewater purification by Chlorella vulgaris. Biodegradation. 24: 721-732.

28. Yoshihara, K. I., Nagase, H., Eguchi, K., Hirata, K., Miyamoto, K. 1996. Biological elimination of nitric oxide and carbon dioxide from flue gas by marine microalga NOA-113 cultivated in a long tubular photobioreactor. J. Ferment. Bioeng. 82-4: 351-354.

29. Zeiler, K. G., Heacox, D. A., Toon, S. T., Kadam, K. L., Brown, L. M. 1995. The use of microalgae for assimilation and utilization of carbon dioxide from fossil fuel-fired power plant flue gas. Energy Convers. Mgmt. 36-6-9: 707-712.

Page 171: Effects of Light, Co2 and Reactor Design on Growth of Algae

157 

 

 

CHAPTER 9 BIOMASS PRODUCTION OF TWO MARINE AND TWO FRESH

WATER ALGAE GROWING IN AN INDUSTRIAL-SCALE THIN FLAT PANEL PHOTO-BIOREACTOR UNDER FOUR CARBON DIOXIDE CONCENTRATIONS

 

9.1 ABSTRACT

Two marine and two fresh water algae were grown in 750 L flat panel photo-bioreactors under

four CO2 concentrations, 395 ppm, 1500 ppm, 4000 ppm and 10,000 ppm, to evaluate CO2

concentration effects on growth rate, population density and biomass production. The fresh water

alga, Chlorella vulgaris reached an average standing biomass of 2582 mg L-1 when the CO2

concentration was 10,000 ppm, and 760 mg L-1 growing at 380-410 ppm of CO2. The algae

Scenedesmus sp increased its biomass from 424 to 2428 mg L-1 with CO2 concentrations

increasing from 395 to 10,000 ppm. The marine alga Nannochloris sp increased its biomass from

594 to 1796 mg L-1 as the CO2 concentration raised from 395 to 10,000 ppm, and Tetraselmis sp

changed its average standing biomass from 710 to 1115 mg L-1 when CO2 changed from 380-410

to 10,000 ppm. The trend of increasing in biomass with increasing CO2 concentrations was tested

with ANOVA repeated measurements resulting in statistical difference in three of the four algae.

Keywords: Biomass production, population density, Carbon dioxide uptake, artificial lighting.

9.2 INTRODUCTION

Flat panel photo-bioreactors (FPB) represent the latest evolution in technology for industrial-

scale, engineered systems for production of algal biomass. Biomass production under these

systems greatly exceeds those achieved in open or closed ponds, and in closed bubble systems or

closed tubular photo-bioreactors. The main advantage of FPB technology is the thickness of the

system; a thin layer (5-70 mm) of growing medium which increases light exposure to the algae

allowing it’s photosynthetic machinery to operate more efficiently. Another advantage of the

Page 172: Effects of Light, Co2 and Reactor Design on Growth of Algae

158 

 

technology is the higher efficiency of oxygen release during photosynthesis, thereby reducing

oxidative damage to the photosynthetic machinery (Sierra et al., 2008). Increases in biomass

production from 0.34 to 4.4 g L-1 d-1 have been reported for FPB thickness decreases from 104

mm to 13 mm (Hu et al., 1996).), standing biomass of 16 g L-1 was attained in a FPB with a

thickness of 13 mm, biomass decreased below 3.8 g L-1 when the thickness increased to 52 mm

growing the marine algal Nannochloropsis sp (Zou and Richmond, 1999).

Open ponds have a useful media thickness about of 15-30 cm as light of appropriate intensity is

not able to reach the algae population below these depths. Productivity of open pond systems is

typically 0.06-0.1 g L-1 d-1 with a standing biomass of 1 g L-1 (Pulz, 2001; Becker, 1993). Open

pond systems are also subject to contamination (Pulz, 2001). In bubble systems and closed

tubular photo-bioreactors the algal medium limits the light intensity reaching the growing algae,

however there have been major improvements in biomass production compared to those from

open ponds because in closed systems the risk of contamination is reduced and biomass

production is increased. Further, the functioning cost of closed photo-bioreactors is 43%

compared to 57% of open ponds and the financial returns of photo-bioreactors are 3% more

(Richardson et al., 2012).

A FPB of 6-7 mm thickness can achieve a biomass production of 5.6 g L-1 d-1 and standing

biomass of 40 g L-1 (Doucha and Livansky, 2009). A thin layer reactor 10 mm thick achieved a

standing biomass of 10 g L-1 and biomass production of 1.7 g L-1 (Livasky and Doucha, 1997).

With a FPB of 200 mm thickness, a specific growth rate of 0.52 d-1 has been reported with the

alga Haematococcus pluvialis (Issarapayup et al., 2009). The efficiency of a FPB of 6 mm

thickness reached 49% CO2 absorption with a net efficiency of 4.4 kg of CO2 to produce 1 kg of

dry biomass. This represents a net absorption efficiency of 38.7% of the supplied CO2; 10.3% is

absorbed and lost from medium and 51% is not absorbed at all (Doucha et al., 2005).

Page 173: Effects of Light, Co2 and Reactor Design on Growth of Algae

159 

 

Though FPB have a higher efficiency in biomass production and carbon uptake efficiency per

medium volume than open or closed pond and tubular photo-bioreactors, they have a lower

medium volume per square meter bioreactor area. Further, the cost of this technology is much

higher on an area and volume basis than other technologies. This study examined algal growth in

a FPB which addressed the volume/area problem by positioning the bioreactor vertically.

9.3 MATERIALS AND METHODS

9.3.1 MICROORGANISM

The four algal species grown this study are from the culture collection kept at the University of

Texas, Austin, TX; Chlorella vulgaris (26), Nannochloris sp (1268), Scenedesmus sp (1589) and

Tetraselmis sp (2767).

9.3.2 FLAT PANEL REACTOR

The FPB designed and constructed for this study contained 8 units, each unit consisting of 15 flat

panels constructed from polypropylene and connected to a 500 L tank. Each flat panel was

located 35-40 cm between one another, which simulated a 30-m tube, was comprised of 15

channels, each 2 m in length, 2.4 m in height and 30-50 mm in width. The entire unit contained

750 L of algal medium. A submersible pump (500 watt) set in the tank was capable of circulating

medium at 7,000 -7,500 L h-1. Each flat panel bioreactor held 30-33 L of medium with a flow

rate of 7-8 L-1 min-1 reactor-1 (the flow of medium in the reactor was 10-12 cm sec-1). The

distribution of the algal media was 250-260L resting in the 500L tank and 450-495L moving in

the 15 flat panel photo bioreactors.

A pump was used raise the medium to the top of the reactor and it flowed down through the

under gravity. The exposure time of the algal medium was 4 min per flow through the reactor

and in total there were 8-10 complete medium cycles per hour, considering the time that the

medium was in the tank before being pumped up to the top of the reactor. Figure 9.1 shows the

FPB set up.

Page 174: Effects of Light, Co2 and Reactor Design on Growth of Algae

160 

 

Figure 9.1. Reactor unit from experimental set up.

9.3.3 ILLUMINATION AND TEMPERATURE

During daytime the system was exposed to solar radiation, the polyethylene-covered greenhouse

allowing 90% of the total solar radiation to reach the bioreactors. A portion of the solar UV

radiation was filtered by the greenhouse cover and by the polypropylene panels, the solar

radiation reaching the reactors through each day ranged from 71 to 1500 µmol m-2 s-1, outside the

greenhouse the maximum radiation registered was 2550 µmol m-2 s-1, measured by a quantum

meter (Model MQ, Apogee Instruments Inc, Logan, UT, USA). During the night time artificial

illumination was provided to each reactor by a fluorescent cold white lamp (180 watts). In

addition there were two blue (425 nm) 52 watt reflectors made up of LEDs and supplying 2,800

µmol m-2 s-1, measured 5 cm from the reflector. In total 1544 watts were supplied to 6, 000 L of

medium (750 L * 8 reactors = 6000 L) for an average input of 0.257 W L-1, See Figure 9.2 for

night illumination. Temperature in the media was 23°C + 2 during the entire time of the

experiments. Temperature sensors were used to measure temperature outside of the greenhouse,

temperature inside of the greenhouse and temperature of the reactors media at all time.

Page 175: Effects of Light, Co2 and Reactor Design on Growth of Algae

161 

 

Figure 9.2. Experimental set up showing the night illumination.

9.3.4 CARBON DIOXIDE SUPPLY

Carbon dioxide was supplied to the FPB using a air turbine (Model GF-180, Resun, China)

which provided 0.9 m3 min-1 at 11.7 kPa. The turbine combined air from the atmosphere with

pure carbon dioxide from a supply tank, and introduced the air-CO2 mixture into the eight 750 L

reactors. The air-CO2 mixture was measured continuously with a CO2 sensor (Model 7001,

Telaire-General Electric, California, USA). Four CO2 concentrations were supplied: 395 ppm

(atmospheric air), 1,500 ppm, 4,000 ppm and 10,000 ppm CO2.

 

9.3.5 CULTURE MEDIUM

The culture medium was prepared using a commercial synthetic fertilizer, Solucat (25-5-5 +

trace elements), by dissolving 0.8 g in 1 L distilled water. At the start of each experiment the

culture medium contained 114 mg L-1 ammonium-N, 86.4 mg L-1 nitrate-N, 40 mg L-1

phosphate, 40 mg L-1 potash, 160 μg L-1 iron, 80 μg L-1 manganese, 80 μg L-1 boron, 16 μg L-1

zinc, and 16 μg L-1 copper; the electrical conductivity was 0.95 ± 0.01 mS, measured by EC

Page 176: Effects of Light, Co2 and Reactor Design on Growth of Algae

162 

 

sensor (Model HI 991301, Hanna, Michigan, USA). This EC level was maintained throughout

the experiment. For the two marine algal species, NaCl was added to the medium to increase the

salt concentration to 3.5%, simulating marine conditions.

9.3.6 BIOMASS DETERMINATION

A colorimetric methodology was used to estimate production of algal biomass. A calibration

curve was prepared by measurement of the absorbance at 680 nm (Shimadzu UV-1600,

Shimadzu Scientific Instruments Inc, Columbia, USA), for 7 known algal biomass ranging from

0.17-3.49 g L-1 in suspension medium. Linear regression analysis of the data for each alga gave

the following equations: for Nannochloris sp, Y= 0.8586x + 0.1464 with R2 = 0.9854, for

Tetraselmis sp, Y= 1.054x + 0.2538 with a R2 = 0.9985, for Scenedesmus sp Y= 0.81x + 0.0461

with R2 = 0.9813, and for Chlorella vulgaris, Y= 0.7015x + 0.0362 with R2 = 0.9887.

We used the following methodology to measure the biomass: (1) shake the 15 flat panel in each

unit to suspend algae growing in the walls of the flat panel, (2) wait 3-5 minutes to ensure that

the system is homogenous, (3) take a sample from the nutrient supply tanks with a 20 mL

syringe, (4) introduce the sample into a spectrophotometer vial, (5) take the absorbance reading

at 680 nm, (6) repeat steps 4 to 5 and compare results to ensure that they are similar and take the

average of the two results, (7) in the case of different in the results repeat the six steps before.

9.3.7 EXPERIMENTAL DESIGN AND STATISTICAL ANALYSIS

The experiment measured population density and standing biomass of four algae species

supplied with four different CO2 concentrations over an18 d incubation. The algae were assigned

to the reactors at random. Each algal treatment was replicated twice, with the 8 reactor units

receiving the same concentration of CO2 concentration during the incubation. The results were

analyzed using a one and two way ANOVA and repeated measures ANOVA using the software

Biostatistics 1.0 and Stats Pad version 1.3.

Page 177: Effects of Light, Co2 and Reactor Design on Growth of Algae

163 

 

9.4 RESULTS

9.4.1 CHANGES IN ALGAL CELL DENSITY UNDER AIR SUPPLY (380-410 PPM OF CO2)

The four algae species showed a typical sigmoid growth curve, confirming that the industrial flat

panel system was sufficiently well-designed to support algae growth under these conditions

(Figure 9.3). The marine algae Tetraselmis sp developed the highest population density

3.58*1007 cells mL-1; the marine algal Nannochloris sp produced a population density of

2.44*1007 cells mL-1. The two fresh water species produced less cell density, though their cell

sizes are larger. The cell density of Scenedesmus sp reached 1.04*1007 cells mL-1 whereas C.

vulgaris developed the lowest population density 1.31*1007 cells mL-1, though it produced a

higher biomass.

Figure 9.3. Population density of four alga species growing in a flat panel industrial scale reactor.  

Page 178: Effects of Light, Co2 and Reactor Design on Growth of Algae

164 

 

Growth rate of the fresh water algae C. vulgaris was 1.47 in 18 d-1 and for Scenedesmus sp was

1.45 in 18 d-1; on the marine algae growth rate for Tetraselmis sp was 1.36 in 18 d-1 and for

Nannochloris sp was 1.43 in 18 d-1.

9.4.2 BIOMASS PRODUCTION WITH INCREASING CO2 TREATMENTS

Figure 9.4 shows the substantial increase in algal biomass production after an 18-d incubation

when exposed to increasing CO2 concentrations. This trend is most evident in the two fresh water

species: C. vulgaris increased its biomass 3.4-fold, from 760 mg L-1 to 2582 mg L-1, whereas

Scenedesmus sp increased its biomass from 424 mg L-1 to 2428 mg L-1, a 5.7 fold increase. The

marine algae Nannochloris sp increased its biomass from 691 mg L-1 to 1796 mg L-1 and

Scenedesmus sp increased its biomass from 710 mg L-1 to 1115 mg L-1.

Figure 9.4. Biomass production of two fresh water algae C. vulgaris, Scenedesmus sp and two marine algae Tetraselmis sp, Nannochloris sp exposed to increasing CO2 concentrations in an FPB.

Page 179: Effects of Light, Co2 and Reactor Design on Growth of Algae

165 

 

 

9.4.3 BIOMASS COMPARISON OF ALGAE GROWING AT DIFFERENT CO2

CONCENTRATIONS

The fresh water algal C. vulgaris in general produced more biomass than the other algal species.

At a CO2 concentration of 380 - 410 ppm, C. vulgaris produced a standing biomass of 761 mg L-

1 with a peak of 1009 mg L-1, which is similar to Tetraselmis sp at 766 mg L-1 with a peak of 867

mg L-1. When the CO2 concentration raised to 1000-2000 ppm C. vulgaris biomass was 1284 mg

L-1 with a peak of 1946 mg L-1 and the highest standing biomass was from Nannochloris sp 1330

mg L-1 with a peak of 2038 mg L-1. C. vulgaris produced the highest standing biomass 1741 mg

L-1 when the CO2 concentration was 4000 ppm with a peak of 2067 mg L-1. When the CO2

concentration raised to 10,000 ppm the standing biomass of C. vulgaris was 2316 mg L-1 with a

peak of 3057 mg L-1 and the highest average standing biomass was 2357 mg L-1 produced by

Scenedesmus sp with a peak of 2665 mg L-1. Consistently C. vulgaris produced the highest or

second highest standing biomass at all CO2 concentration treatments.

A two way ANOVA test was made to compare four different algae and four different CO2

concentration, resulting in statistical differences with F in the samples = 19.4936, columns =

63.5129, and interaction = 6.38448 and p = 0.000. In addition the two variables; algae species

and CO2 concentration, were analysed separately. An ANOVA repeated measurements showed

that there were statistical differences in biomass production with increasing CO2 concentrations

in C. vulgaris F= 24.5948 with p =0.0130, in Scenedesmus sp F = 123.1045 with p = 0.0012, in

Nannochloris sp F = 23.1414 with p = 0.0141, though it was not significantly different in

Tetraselmis sp F = 3.4339 with p = 0.1690. Therefore, increases in CO2 concentration increased

algal biomass production significantly when the changes are from 400 ppm to 10,000 ppm of

CO2 in three of the four species of algae tested under the present research.

A one way ANOVA was performed to compare the biomass of the four algae species at specific

CO2 concentration; for the four algae growing at 380-400 ppm the results were F = 5.0169 with p

= 0.0058, at 1000-2000 ppm F = 5.0732 with p = 0.0055, at 4000 ppm of F = 22.7575 with p =

0.0000 and for 10,000 ppm we got F = 43.2615 with p = 0.0000. As a result of these tests it was

Page 180: Effects of Light, Co2 and Reactor Design on Growth of Algae

166 

 

found statistical differences in the biomass production from algae to algae growing at the same

CO2 concentration, therefore each alga produced difference quantity of biomass growing at the

same time in equal flat panel reactors at equal CO2 concentration.

9.5 DISCUSSION AND CONCLUSION

The biomass production reached by the four algae species in an 18-d incubation growing in an

FPB with thickness of 30-50 mm were from 424 mg L-1 to 2428 mg L-1. This biomass is similar

to those obtained in previous studies. A study conducted with C. vulgaris reached a standing

biomass of 1700 mg L-1 after 12-15 days of incubation (Liang et al., 2009). A study with C.

vulgaris growing in an FPB of 15 mm and 30 mm thickness generated a standing biomass of

1950 mgL-1 and 1800 mgL-1 respectively (Degen et al., 2001). A most recent study resulted in

more standing biomass 900 mg L-1 a lower reactor thickness 50-100 mm after 30 days

incubation, and a maximum standing biomass of 4640 mg L-1 after 100 days incubation (Yoo et

al., 2013). Another study growing C. vulgaris in a FPB of 3 mm reached 12450 mg L-1 in just 11

days of incubation (Choi et al., 2013). In the present study, C. vulgaris produced more biomass

than the other three algae species growing under to similar CO2 concentrations.

Although the FPB is a reliable system to grow algae for high biomass production in industrial

scale, more studies are needed to developed its maximum potential in biomass production per

volume and to increase the volume-area ratio.

ACKNOWLEDGMENTS

This research was funded by the Government of Canada through the grants of Professor Paul

Voroney at the School of Enviromental Sciences, University of Guelph, Ontario Canada.

Page 181: Effects of Light, Co2 and Reactor Design on Growth of Algae

167 

 

9.6 REFERENCES

1. Becker E. W. 1993. Microalgae Biotechnology and Microbiology. Cambridge University Press, Cambridge 304p.

 

2. Choi, H.J., Lee, H.M., Lee, S.M. 2013. A novel optical panel photobioreactor for cultivation of microalgae. Water. Sci. tech. 67-11: 2543-2548.

3. Degen, J., Uebele, A., Retze, A., Schmid-Staiger, U., Trosch, W. 2001. A novel airlift photobioreactor with baffles for improved light utilization through the flashing light effect. J. Biotechnol. 92: 89-94.

4. Doucha, J., Livansky, K. 2009. Outdor open thin-layer microalgal photobioreactor: potential productivity. J. Appl. Phycol. 21: 111-117.

5. Doucha, J., Straka, F., Livansky, K. 2005. Utilization of flue gas for cultivation of microalgae Chlorella sp in an outdor open thin-layer photobioreactor. J. Appl. Phycol. 17: 403-412.

6. Hu, Q., Guterman, H., Richmond, A. 1996. A flat inclined modular photobioreactor for outdoor mass cultivation of photoautotrophs. Biotechnol. Bioeng. 51: 51-60.

7. Issarapayup, K., Powtongsook, S., Pavasant, P. 2009. Flat panel airlift photobioreactors for cultivation of vegetative cells of microalga Haematococcus pluvialis. J. Biotechnol. 142: 227-232.

8. Liang, Y; Sarkany, N; Cui, Y. 2009. Biomass and lipid productivities of Chlorella vulgaris under autotrophic, heterotrophic and mixotrophic growth conditions. Biotechnol. Lett. 31: 1043-1049.

Page 182: Effects of Light, Co2 and Reactor Design on Growth of Algae

168 

 

9. Livansky, K., Doucha, J. 1997. Additional CO2 saturation of thin-layer outdoor microalgal cultures: CO2 mass transfer and absorption efficiency. Algological Studies. 87: 145-154.

10. Pulz, O. 2001. Photobioreactors: production systems for phototrophic microorganisms. Appl Microbiol. Biotechnol. 57: 287-293.

11. Richardson, J. W., Johnson, M. D., Outlaw, J. L. 2012. Economic comparison of open pond raceways to photo bio-reactors for profitable production of algae for transportation fuels in the southwest. Algal Research. 1: 93-100.

12. Sierra, E., Acien, F. G., Fernandez, J. M., Garcia, J. L., Gonzalez, C., Molina, E. 2008. Characterization of a flat plate photobioreactor for the production of microalgae. Chem. Eng. J. 138: 136-147.

13. Yoo, J.J., Choi, S.P., Kim, J.Y.H., Chang, W.S., Sim, S.J. 2013. Development of thin-film photo-bioreactor and its application to outdoor culture of microalgae. Bioprocess. Biosyst. Eng. 36: 729-736.

14. Zou, N., Richmond, A. 1999. Effect of light-path length in outdoor flat plate reactors on output rate of cell mass and of EPA in Nannochloropsis sp. J. Biotechnol. 70: 351-356.

Page 183: Effects of Light, Co2 and Reactor Design on Growth of Algae

169 

 

 

CHAPTER 10: GENERAL CONCLUSIONS

During the course of this research, we evaluated biomass production and population density of

marine and fresh water algae growing under different parameters such as light quality, CO2

concentration and different types of photo-bioreactors. For light treatments, 6 algae species, three

marine and three fresh water algae, were grown under solar radiation, blue-red LEDs and

fluorescent illumination. The alga Chlorella vulgaris was grown under four different LED lights,

blue, red, white and blue-red, and one marine and one fresh water algae were grown under

flashing LED lights. For CO2 concentration we grew the algae Chlorella vulgaris under four

different CO2 concentration and four different light treatments. We grew two marine and two

fresh water algae under a polypropylene flat panel reactor and measured the biomass production

of the algae for 5 months under four different CO2 concentrations.

All light treatments that included solar radiation showed a faster growth rate and higher biomass

production than those experiments that did not include solar radiation and relied only on artificial

illumination. The treatments that produced the highest biomass were obtained when we

combined solar radiation during the day with artificial illumination during the night period. In

general, these treatments also performed better than those treatments with just 24 h artificial

illumination. The experiments which compared different wavelengths resulted in more biomass

production when the algae Chlorella vulgaris grew under white LED, the cell weight was higher

under white light illumination, and the one way ANOVA test confirmed these results. The

second best treatment was the combination of red-blue LEDs, which produced the second highest

biomass production and the second highest cell weight. The flashing light experiment for a

marine alga Nannochloris sp and fresh water alga Scenedesmus sp showed more standing

biomass and higher cell weights in both algae under flashing lights compared against continuous

light. There was a statistical difference in daily biomass production and growth rate under

flashing light against continuous light in the marine alga Nannochloris sp, however there was not

Page 184: Effects of Light, Co2 and Reactor Design on Growth of Algae

170 

 

conclusive evidence from this study for the fresh water alga Scenedesmus sp because the

ANOVA test resulted in no significant difference in daily biomass production and growth rate.

The experiments conducted with the alga Chlorella vulgaris grown under different

concentrations of CO2 resulted in more standing biomass with increasing CO2 concentrations.

We also tested different light treatments and confirmed that Chlorella vulgaris growth was better

under white LED and Blue LED. The ANOVA test confirms the fact of more algal standing

biomass as the concentration of CO2 in the inlet air rise from 0.035% to 9%.

We tested for five months a new flat panel photo-bioreactor with two marine and two fresh water

algae. These industrial scale experiments confirm what we previously study in Chlorella

vulgaris, more biomass production as the CO2 concentration in the inlet air raise. The two marine

algae tested Nannochloris sp, Tetraselmis sp and the two fresh water algae Chlorella vulgaris,

Scenedesmus sp increased their biomass production when the CO2 in the inlet air increased from

380 ppm, 1000-2000 ppm, 4000 ppm and 10,000 ppm, thus the ANOVA test confirm this trend.

The flat panel photo-bioreactor produced more biomass using Chlorella vulgaris at 1% of CO2

than some of the previous experiments using the bubble column photo-bioreactor at 8-9% of CO2

concentration, thus the standing biomass obtained growing the two marine and the two fresh

water algae in the flat panel photo-bioreactor were the highest during the course of this study.

During the course of our experiments we worked with red (625 nm) illumination and this light

treatment showed the worst performance in all experiments.The algae has an absorption peak at

around 660 nm and the mean chlorophyll reaction centers are 670 and 680 nm. For this reason

we suggest more studies comparing blue and white lights against red 660-690 nm, and all three

treatments together. These experiments would provide conclusions on the effects of different

light wavelengths on algal biomass production for selection of the best light source or

combination of lights. We also suggest that more experiments should be conducted with a

flashing light source because this technology decreases energy consumption up to 90% and

increases growth rate and biomass production. The research with CO2 concentrations should be

Page 185: Effects of Light, Co2 and Reactor Design on Growth of Algae

171 

 

expanded to evaluate increasing the concentration to 15, 20, 25, 30%, and higher concentrations

of CO2 to determine the optimal concentration for algal growth and for highest CO2 removal rate.

The experiments with the flat panel reactor showed the need for further research to increase

biomass production and robustness of the technology for industrial scale. However biomass

production in the flat panel reactor is sufficient for commercial production of algae. Algal

biomass can be used as a component of feedstocks for fish, cattle and pigs. The oils contained in

the algae, such as omegas 3, 6 and 9, are valuable products for human food supplement. The fast

growing alga C vulgaris is rich in high value molecules such as lutein and canthaxanthin. These

two powerful antioxidants are currently marketed as human supplements, making these algal

species attractive for use as a crop.

In general the six algal species tested in this study contain oils that can be transformed into

biodiesel and the carbohydrates can be fermented for commercial ethanol production. The ability

of algae to rapidly fix high quantities of CO2, plus the importance of the constituents of the algae

such as proteins, high value molecules for food supplements, fats and carbohydrates for biofuels,

make algae an alternative for a new generation of crops. They can be established on land that is

unsuitable for agriculture, such as deserts, and grown using contaminated water or sea water.

This research has obtained interesting results, some conclusive and some not, althought there is

potential to further increase biomass production efficiencies. The biomass production obtained is

enough to sustain commercial algal production. Critical factors such as illumination intensity and

composition, CO2 concentration in the growing media and reactor design have an impact on

biomass production. The flat panel reactor at 8% CO2 air intake, and the combination of solar

radiation and multiple LEDs lights (red, blue and white) with the flashing effect during the night

period maximizes biomass production. The combination of these critical factors is the most

efficient way to grow algae from the findings of this research.