113
Erasmus Mundus Master Course: IMETE Thesis submitted in partial fulfilment of the requirements for the joint academic degree of: International Master of Science in Environmental Technology and Engineering an Erasmus Mundus Master Course from Ghent University (Belgium), ICTP (Czech Republic), UNESCO-IHE (the Netherlands) Nitrification and denitrification by algal-bacterial biomass in a Sequential Batch Photo-bioreactor: effect of SRT Host university: MSc Thesis by Dudy Fredy Supervisor Mentor Prof. Piet Lens Dr. Peter van der Steen Delft August 2013 This thesis was elaborated and defended at the UNESCO-IHE, Delft, The Netherlands within the framework of the European Erasmus Mundus Programme “Erasmus Mundus International Master of Science in Environmental Technology and Engineering " (Course N° 2011-0172)

Nitrification and denitrification by algal-bacterial

  • Upload
    others

  • View
    5

  • Download
    0

Embed Size (px)

Citation preview

Erasmus Mundus Master Course: IMETE

Thesis submitted in partial fulfilment of the requirements for the joint academic degree of:

International Master of Science in Environmental Technology and Engineering

an Erasmus Mundus Master Course from

Ghent University (Belgium), ICTP (Czech Republic), UNESCO-IHE (the Netherlands)

Nitrification and denitrification by algal-bacterial biomass in a Sequential Batch Photo-bioreactor:

effect of SRT

Host university:

MSc Thesis by

Dudy Fredy

Supervisor Mentor

Prof. Piet Lens Dr. Peter van der Steen

Delft August 2013

This thesis was elaborated and defended at the UNESCO-IHE, Delft, The Netherlands within the

framework of the European Erasmus Mundus Programme “Erasmus Mundus International Master of

Science in Environmental Technology and Engineering " (Course N° 2011-0172)

Certification

This is an unpublished MSc. thesis and is not prepared for further distribution. The author and the

promoter give the permission to use this thesis for consultation and to copy parts of it for personal use.

Every other use is subject to copyright laws, more specifically the source must be extensively

specified when using results from this thesis.

The Promoter The Author

Prof. Piet Lens

Dudy Fredy

The findings, interpretations and conclusions expressed in this study do neither necessarily reflect the

views of the UNESCO-IHE Institute for Water Education, nor of the individual members of the MSc

committee, nor of their respective employers.

for Azzania and Dyaz

i

Abstract

A post-treatment of UASB reactor’s effluent by utilizing the interaction of algae and bacteria can offer

lower energy consumption via photosynthetic oxygenation. To meet a stringent standard effluent it is

necessary to optimize nitrogen removal. This research investigated the effects of Sludge Retention

Time (SRT) on nitrification and denitrification performance.

A mixed biomass culture of different species of microalgae and bacteria were inoculated in an open

photo-bioreactor. The 1-L reactor was illuminated (25.9 µmol/m2.s) and operated as a sequential

batch reactor at 28ºC and pH 7.5 with 50% discharge per cycle (2 cycles per day). SRT were varied

from 17 to 52 days by discharging different volume of completely mixed culture.

Nitrification and denitrification of artificial wastewater (23 mgN-NH4/L and 200 mg COD/L) were

achieved. The present study identified that SRT has only limited effect on the nitrification process and

uptake by algal-bacterial biomass in a photo-bioreactor. The overall removal rates only varies from 2.1

to 2.9 mgN-NH4/L.h , while SRT varies from 17 to 52 days. The maximum oxygen production was

occurred in SRT 17 days at a rate 0.3 mgO2/m3.day.

Keywords: algae, denitrification, nitrification, photo-bioreactor, photosynthetic oxygenation, SRT

iii

Acknowledgements

I would like to thank the almighty Allah for the uncountable blessing that He has bestowed upon me.

My sincerest appreciations go to Dr. Peter van der Steen. Thank you so much for your guidance, input,

throughout the research and thesis writing period. I am deeply grateful to Prof. Piet Lens for the advice

and input throughout thesis work.

Thank you to Angelica Rada and Rudatin Windaswara, for countless support, input and of course the

valuable data for completing this thesis.

I acknowledge Eldon Raj, Carlos Lopez Vazquez, and all the colleagues in paper writing meeting for

correction and suggestions towards the finalizing of this thesis. I also thank to my colleagues in algae

research: Indri Karya, Kuntarini Rahsilawati, Freweyni Tammene, and Thanh Tung Nguyen.

I would like to express my appreciation for the help and assistant of the laboratory staff at UNESCO-

IHE: Fred Kruis, Berend Lolkema, Peter Heering, Ferdi Battes, Frank Wiegman and Lyzette

Robbemont. I also would like to thanks to Laurens Welles, Javier Sánchez Guillén and Sondos Saad

who were very helpful during ‘great’ months in the lab.

The last and the most important, my wife Azzania. All is nothing without your support. Thank you for

your pray, understanding and sacrifices. And to my son Dyaz. You are my spirit keeper. Thank you.

v

Table of Contents

ABSTRACT ................................................................................................................................... I

ACKNOWLEDGEMENTS ............................................................................................................ III

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

1.1 Background ................................................................................................................ 1

1.2 Problem Statement ..................................................................................................... 2

2 LITERATURE REVIEW ........................................................................................................ 3

2.1 Post-treatment options of UASB reactor’s effluent ................................................... 3

2.2 Biological nitrogen removal from wastewater ........................................................... 4

2.2.1 Nitrification ............................................................................................................ 4

2.2.2 Denitrification ........................................................................................................ 5

2.2.3 Effect of C/N ratio on nitrogen removal ................................................................ 7

2.2.4 SRT (Solids Retention Time) ................................................................................. 7

2.2.5 Sequential Batch Reactor (SBR) ............................................................................ 8

2.3 Algal-bacterial consortium for wastewater treatment .............................................. 10

2.3.1 Microalgae ............................................................................................................ 13

2.3.2 Microalgae culture system ................................................................................... 13

2.3.3 Influence of environmental parameters on algal growth ...................................... 14

2.4 Effect of dark periode on nitrogen removal ............................................................. 17

3 OBJECTIVES ..................................................................................................................... 19

3.1 General objective ...................................................................................................... 19

3.2 Specific objectives .................................................................................................... 19

4 MATERIAL AND METHODS .............................................................................................. 21

4.1 Culture medium ........................................................................................................ 21

4.2 Microalgae-bacteria consortium ............................................................................... 22

4.3 Reactor set up and experimental design ................................................................... 22

4.4 Sampling ................................................................................................................... 24

4.5 Analytical Methods .................................................................................................. 24

4.5.1 Ammonium nitrogen ............................................................................................ 24

4.5.2 Nitrite nitrogen ..................................................................................................... 25

4.5.3 Nitrate nitrogen .................................................................................................... 25

4.5.4 Chlorophyll-a ....................................................................................................... 25

4.5.5 Total suspended solid (TSS) and volatile suspended solid (VSS) ....................... 26

4.5.6 Biomass Light Absorption ................................................................................... 26

4.6 Calculation ............................................................................................................... 26

4.6.1 Solid Retention Time (SRT) ................................................................................ 26

4.6.2 Nitrogen balance .................................................................................................. 27

4.6.3 Biomass composition ........................................................................................... 28

4.6.4 Oxygen production by algae ................................................................................ 29

4.7 Statistical analysis .................................................................................................... 31

4.8 Batch experiment ...................................................................................................... 31

5 RESULTS ........................................................................................................................... 33

5.1 Nitrogen removal ...................................................................................................... 33

5.1.1 Daily nitrogenous concentration .......................................................................... 33

5.1.2 Ammonium conversion rate ................................................................................. 34

5.2 Nitrogen balance ...................................................................................................... 35

5.3 Chlorophyll-a concentration ..................................................................................... 36

5.4 Suspended solids concentration ............................................................................... 37

5.5 Light absorption ....................................................................................................... 38

5.6 Biomass composition ............................................................................................... 39

5.7 Solids Retention Time (SRT) ................................................................................... 39

5.8 Oxygen production ................................................................................................... 41

5.9 Nitrate uptake batch experiment .............................................................................. 44

5.10 Microscopic observation .......................................................................................... 44

6 DISCUSSION ...................................................................................................................... 47

6.1 Ammonium conversion rate ..................................................................................... 47

6.2 The effects of different operational sequences in SBR ............................................ 48

6.3 The effects of different SRT in SBR ........................................................................ 49

6.4 Comparison with other algal-bacterial photo-bioreactor ......................................... 51

6.5 Development of biofilm in the reactor ..................................................................... 52

6.6 Light regime in photobioractor ................................................................................ 53

6.7 Denitrification .......................................................................................................... 54

7 CONCLUSION AND RECOMMENDATIONS ......................................................................... 55

7.1 Conclusion ................................................................................................................ 55

7.2 Recommendations .................................................................................................... 55

REFERENCES ............................................................................................................................ 57

APPENDIX A ............................................................................................................................. 63

APPENDIX B ............................................................................................................................. 73

APPENDIX C ............................................................................................................................. 84

APPENDIX D ............................................................................................................................. 92

APPENDIX E ............................................................................................................................. 93

vii

List of tables

Table 2.1: Typical average concentration of UASB reactor’s effluent ................................................... 3

Table 2.2: Summary of various UASB-post treatment systems and their average level of effluent

quality ................................................................................................................................. 3

Table 2.3: Different operational sequences of SBR operation ................................................................ 9

Table 2.4: Algal-bacterial consortium for nutrient removal from wastewater ...................................... 13

Table 4.1: Modified BG-11 medium for microalgae and bacteria culture ............................................ 21

Table 4.2: Experimental variations ....................................................................................................... 23

Table 4.3: SBR operational setting at period 1 ...................................................................................... 23

Table 4.4: SBR operational setting at period 2, 3 and 4 ........................................................................ 24

Table 5.1: Nitrogen removal efficiency................................................................................................. 33

Table 5.2: Summary of ammonium conversion rate in different period ............................................... 34

Table 5.3: Nitrogen balance based on one cycle operation in SBR ...................................................... 35

Table 5.4: Actual SRT calculation ........................................................................................................ 39

Table 5.5: Estimation of biomass composition ..................................................................................... 40

Table 5.6: Estimation of oxygen production and consumption rate in the reactor ................................ 43

Table 6.1: Comparison the result with other research ........................................................................... 51

ix

List of figures

Figure 1.1:Typical configuration of water treatment plant with UASB-Activated Sludge ..................... 1

Figure 2.1:Typical cycles in Sequential Batch Reactor ........................................................................... 9

Figure 2.2:Interaction between algae and bacteria in wastewater treatment process ............................ 11

Figure 2.3: Relation between light intensity on photoautotrophic growth of photosynthetic cells ....... 14

Figure 4.1: Schematic diagram of reactor set up ................................................................................... 22

Figure 4.2: DO concentrations profile partition in one cycle of SBR operation ................................... 29

Figure 5.1: Daily nitrogenous concentration ......................................................................................... 33

Figure 5.2: Ammonium conversion in one cycle of SBR operation in period 2 day 89 ....................... 34

Figure 5.3: Profile of chlorophyll-a concentration in different periods ................................................ 36

Figure 5.4: Average chlorophyll-a concentration in different periods .................................................. 36

Figure 5.5: Profile of SS concentration in different periods.................................................................. 37

Figure 5.6: Average SS concentration in different periods ................................................................... 37

Figure 5.7: Light intensity measurement points (a) and a simplified side view of the reactor (b) ........ 38

Figure 5.8: Estimated light penetration inside reactor ........................................................................... 38

Figure 5.9: Biomass compositions in each period ................................................................................. 39

Figure 5.10: Typical DO profile in one cycle operation in period 1(day 46) ........................................ 41

Figure 5.11: Typical DO profile in one cycle operation in period 2(day 68) ........................................ 42

Figure 5.12: Typical DO profile in one cycle operation in period 3 (day 117) ..................................... 42

Figure 5.13: Typical DO profile in one cycle operation in period 4 (day 171) ..................................... 43

Figure 5.14: Nitrate uptake performances by algae-bacteria consortium .............................................. 44

Figure 5.15: Chlorella sp. and Spirulina sp. (20x magnification) ......................................................... 44

Figure 5.16: Scnedesmus sp. and Anabaena sp. (40x magnification) ................................................... 45

Figure 5.17: Algal-bacterial flocs .......................................................................................................... 45

Figure 6.1: Volumetric productivity of a photobioreactor rUx as a function of biomass concentration Cx

.......................................................................................................................................... 50

Figure 6.2: Attached thread-former species of microalgae in reactor’s wall ........................................ 53

Figure 6.3: Light fraction as function of the chlorophyll-a concentration in the reactor ...................... 53

Figure 6.4: Typical nitrogen concentration profile within a cycle (Day 117) ....................................... 54

xi

Abbreviations and symbols

AOB Ammonia-oxidizing Bacteria

BOD Biochemical Oxygen Demand

BNR Biological Nitrogen Removal

C/N Carbon/Nitrogen

Chla Chlorophyll a

COD Chemical Oxygen Demand

DO Dissolved Oxygen

FA Free Ammonia

FNA Free Nitrous Acid

FSA Free Saline Ammonia

HRAP High Rate Algal Ponds

HRT Hydraulic Retention Time

N Nitrogen

NOB Nitrite-oxidizing Bacteria

OD Optical Density

P Phosphorus

PBR Photobioreactors

SBR Sequencing Batch Reactor

SRT Solid(sludge) Retention Time

TKN Total Kjeldahl Nitrogen

TN Total Nitrogen

TP Total Phosphorus

TSS Total Suspended Solid

UASB Upflow Anaerobic Sludge Blanket

VSS Volatile Suspended Solid

WSP Waste Stabilization Ponds

WWTP Wastewater Treatment Plant

Dudy Fredy 1

1 Introduction

1.1 Background

High-rate anaerobic treatment system especially Upflow Anaerobic Sludge Blanket (UASB) reactor is

popularly used for sewage treatment in tropical and developing countries (Seghezzo et al., 1998;

Gomec, 2010). It is a sustainable technology and offers some advantages such as low cost, low energy

consumption and simple operation (Chernicharo, 2006; Khan et al., 2011). However, the effluents of

anaerobic reactor require a post-treatment to meet stringent discharge standards, especially the

Nitrogen concentration in the effluent. Various technological options for further treating the UASB’s

effluent are available to achieve desired effluent quality. One of the promising options is to couple

UASB with Activated Sludge System. A UASB-Activated Sludge system consists of a UASB reactor,

a continues-flow aerated bioreactor and a settler as shown in Figure 1.1. Many studies reported that

coupling UASB with Activated Sludge can achieve high removal of Biochemical Oxygen Demand

(BOD), Chemical Oxygen Demand (COD) and Total Suspended Solids (TSS) (Chernicharo, 2006;

Chong et al., 2012).

Figure 1.1:Typical configuration of water treatment plant with UASB-Activated Sludge

Source: (Chernicharo, 2006)

As a variant of activated sludge system, Sequential Batch Reactor (SBR) can be coupled with UASB

(Chong et al., 2012). In SBR, treatment of wastewater is carried out in various consecutive phases

such as filling, reaction, settling, decant, and idle, within one reactor. SBR adds up the benefit on

smaller footprint requirement.

On the other hand, microalgae have been used to treat wastewater for many years in large ponds and

photo-bioreactors. Microalgae have high affinity for nitrogen and phosphorous, and do not require

much an organic carbon source (Boelee et al., 2012). The use of microalgae for wastewater treatment

2 MSc Thesis

becomes more attractive as wastewater can be a nutrient supply for microalgal biofuel production.

This will bring the biofuel production more economically viable and sustainable (Boelee et al., 2011).

A combination treatment of wastewater, by utilizing the interaction of algae and bacteria can offer

lower energy consumption via photosynthetic aeration. Microalgae provide the oxygen necessary for

aerobic bacteria to biodegrade organic pollutants and ammonium removal, Furthermore, the

application of algal-bacterial system in a Sequential Batch Photo-bioreactor as a post-treatment for

domestic wastewater can show a great potential. It will reduce aeration cost and land requirement,

which are the major problems in developing countries.

1.2 Problem Statement

Nitrification and denitrification are the main pathways for nitrogen removal in wastewater treatment.

In the algal-bacterial system, nitrogen accumulation into algal biomass could also contribute to the

removal process. Previous research (Karya et al., 2013) identified a stable consortium of algae

(Scenedesmus) and nitrifiers that could be established in a photo-bioreactor. It was found that

nitrification could takes place at rates up to 8.5 mg/L.h. The in-situ oxygen generation by algae was

found to be more than sufficient for the nitrification process (Karya et al., 2013).

Beside nitrification, denitrification should also be introduced in the reactor to achieve high quality of

effluent. The removal of nitrogen in the SBR system can be achieved by alternating aerobic and

anoxic periods during the reaction. In photo-bioreactor anoxic periods can be done by creating dark

periods or dark zones. Another study (Windraswara, 2013) identified that denitrification can occur in

algal-bacterial system in a photo-SBR, when no lights was applied in the anoxic periods.

With a more stringent standard effluent for total nitrogen, it is necessary to establish a good nitrogen

removal in Sequential Batch Photo-bioreactor. The nitrogen removal efficiency is influenced by a

number of factors, such as sludge age, temperature, and carbon availability. The focus of this research

therefore is to investigate the effects of sludge age or Sludge Retention Time (SRT) on nitrification

and denitrification performance.

Dudy Fredy 3

2 Literature review

This chapter describes background information of post-treatment options for UASB reactor’s effluent,

conventional biological nitrogen removal, and algal-bacterial interactions in wastewater treatments.

2.1 Post-treatment options of UASB reactor’s effluent

High-rate anaerobic treatment system especially UASB reactor receives great interests for sewage

treatment in tropical and developing countries (Seghezzo et al., 1998; Gomec, 2010). It offers some

advantages such as low cost, low energy consumption and simple operation (Chernicharo, 2006).

However, the effluents of anaerobic reactor require further post-treatment to meet stringent discharge

standards. A typical average effluent concentration from UASB reactor is shown in Tabel 2.1.

Table 2.1: Typical average concentration of UASB reactor’s effluent

Parameter Concentration (mg/L) BOD5 70-100 COD 180-270 TSS 60-100 Ammonia >15 Total N >20 Total P >4

Source: (Chernicharo, 2006)

Many studies on finding the suitable post-treatment have been done. Chong et al., (2012) thoroughly

reviewed the most common options for UASB reactor effluent treatments. The summary of various

UASB-post treatment systems and their effluent quality can be seen in Table 2.2. The main goal of

adopting a post-treatment system to treat an anaerobic effluent is to find a process that is simple in

operation and maintenance, lower capital costs, and energy efficient.

Table 2.2: Summary of various UASB-post treatment systems and their average level of effluent quality

Post-treatment Unit BOD

(mg/L) COD

(mg/L) TSS

(mg/L) Ammonia

(mg/L) TKN

(mg/L)

Total

N

(mg/L)

Total

P

(mg/L) Activated sludge (AS) 7 47 13 3 no data no data no data

Sequencing-batch reactor (SBR) 5 37 12 0.5 10 no data 1.3 Biofilter (BF) 31 70 21 20 20 no data no data

Downflow-hanging sponge (DHS) 8 57 28 13 no data 34 no data

Stabilising pond (SP) 33 110 60 7.6 no data 9.6 2.1 Series Rotating-biological contactor

(RBC)

no data 56 no data 13 8 no data no data

Constructed wetland (CW) 17 47 12 17 36 33 1.9 Source : Adapted from (Chong et al., 2012)

4 MSc Thesis

Based on Table 2.2, the coupling of a post treatment unit to a UASB reactor is effective at removing

the residual organic matter and suspended solids. But, there is still a lack of studies on the removal of

nutrients (Total Nitrogen and Total Phosphorous concentrations). However for achieving a good

nitrogen removal, the organic matter removal efficiency of UASB reactor should be no more than 50–

70%, so that there is enough organic matter for the denitrification step (Chernicharo, 2006).

2.2 Biological nitrogen removal from wastewater

Activated sludge process is one of the most common methods for treating wastewater treatment, in

which nitrification and denitrification are the main pathways for nitrogen removal. Nitrification is a

two steps process carried out by two different autotrophic bacterial groups. In the first step ammonium

is oxidized to nitrite by ammonia oxidizing bacteria (AOB). In the second step nitrite is oxidized to

nitrate by nitrite oxidizing bacteria (NOB). For completing nitrogen removal, nitrate is reduced to

nitrogen gas in the denitrification process by heterotrophic bacteria.

2.2.1 Nitrification

The two sequential oxidation steps in nitrification are shown below as basic stoichiometric redox

reactions:

NH4+ + 1.5O2 NO2

- + H2O + 2H

+ (2.1)

NO2- + 0.5O2 NO3

- (2.2)

By stoichiometri the oxygen requirement for conversion of ammonia to nitrate is 4.57 mgO2/mgN

utilized.

In nitrification process, dissolved oxygen (DO) concentration is the most important parameter for both

AOB and NOB. Ammonia is fully oxidized to nitrate at DO concentrations higher than 1 mgO2/L

(Campos et al., 2007). Higher dissolved oxygen concentrations do not appear to affect nitrification

rates significantly, however low oxygen concentrations reduce the nitrification rate (Ekama and

Wentzel, 2008). Moreover, operation at DO concentrations of 0.6 and 0.4 mgO2/L can cause ammonia

and nitrite accumulations (Campos et al., 2007).

Furthermore, nitrification releases hydrogen ions and decreases alkalinity of the mixed liquor. For

each of mg FSA (Free Saline Ammonia) that is nitrified, 7.14 mg alkalinity (as CaCO3) is consumed.

When the alkalinity falls below about 40 mg/L as CaCO3 then, irrespective of the carbon dioxide

concentration, the pH becomes unstable and decreases to low values (Ekama and Wentzel, 2008).

Hence it is important to maintain a buffer of alkalinity in the aeration tank to provide pH stability and

Dudy Fredy 5

ensure the presence of inorganic carbon for nitrifying bacteria. The residual amount of alkalinity

desired in the aeration tank after complete nitrification is at least 50 mg/L (Gerardi, 2002).

It is reported that the optimum pH for AOB Nitrosomonas is 8.1 and for NOB Nitrobacter is 7.9

(Grunditz and Dalhammar, 2001). The AOB activity is 50% reduced at pH values of 6 and 10

(Jiménez et al., 2012) and NOB were strongly affected by low pH values (no activity was detected at

pH 6.5). And it is also reported that no inhibition was observed at high pH values (activity was nearly

the same for the pH range 7.5–9.95) (Jiménez et al., 2011).

Temperature has a significant effect on microbial nitrification activity. The rate of nitrification is

reduced with decreasing temperature and, conversely, there is a significant acceleration in the rate of

nitrification with increasing temperature. The optimum temperature range for nitrification is between

28-32°C (Gerardi, 2002).

There are some findings in the literature that light might inhibit to both AOB and NOB (Kaplan et al.,

1998; Sinha and Annachhatre, 2006). Illumination with 420 lux (5,000 foot-candles) of light resulted

in complete and irreversible inactivation of ammonia oxidation (Hooper and Terry, 1973). Abelliovich

and Vonshak (1993) reported that light with the intensity of 3.0 x 103 µE/m

2.s had a strong inhibitory

effect on nitrification. On that study the nitrification was done by Nitrosomonas europae in water

containing a high load of organic matter, but not in water with low organic matter (Abeliovich and

Vonshak, 1993). Another study stated also that NOB were more resistance to sunlight than AOB

(Vanzella et al., 1989). However in general, the effect of light depends on the type of nitrifiers as well

as on the environmental condition (Guerrero and Jones, 1996).

2.2.2 Denitrification

Biological denitrification process occurs under anoxic conditions, when the dissolved oxygen

concentration is less than 0.5 mg/L and nitrate ions serve as electron acceptor for microorganisms.

Denitrification can be also termed dissimilatory nitrite/nitrate reduction, because nitrite ions and

nitrate ions, respectively, are reduced to form molecular nitrogen.

There are four sequential steps in denitrification process. The first step is a reduction of nitrate (NO3)

to nitrite (NO2), and followed by second step, where NO2 is reduced to nitric oxide (NO). In the third

step, NO is reduced to nitrous oxide (N2O) an obligate intermediate, some of which ultimately escapes

to the atmosphere. And the last step is the reduction of N2O to nitrogen gas (N2) (Huang et al., 2011).

The overall stoichiometric equation of the conventional denitrification is shown below:

6 MSc Thesis

NO3- + 4gCOD + H

+ 0.5N2 + 1.5g biomass (2.3)

On the contrary to nitrification, denitrification consumes hydrogen ions or generates alkalinity. By

considering nitrate as electron acceptor, it can be shown that for every mg nitrate denitrified, there is

an increase of 3.57 mg alkalinity as CaCO3. Therefore incorporating denitrification in a nitrification

system causes the net loss of alkalinity to be reduced (Ekama and Wentzel, 2008).

The genera Alcaligenes, Bacillus, and Pseudomonas are the largest number of denitrifying bacteria.

Most denitrifiers reduce nitrate ions via nitrite ions to molecular nitrogen without the accumulation of

intermediates. However, some denitrifiers may lack of key enzyme systems to denitrify completely,

then permit the production and accumulation of intermediates (Gerardi, 2002).

Temperature can also affect denitrification rate. The rate is higher with increasing temperature and is

inhibited at wastewater temperature below 5°C. To compensate a decreased rate at low temperature, an

increased Mixed Liquor Volatile Suspended Solid (MLVSS) can enhance the number of denitrifying

bacteria (Gerardi, 2002).

Denitrification can occur over a wide range of pH values. It is reported that denitrification is relatively

insensitive to acidity but may be slowed at low pH (Gerardi, 2002). To ensure acceptable enzymatic

activity of facultative anaerobe and nitrifying bacteria, the pH in the aeration tank should be

maintained at a pH value greater than 7.0. The optimal pH range for denitrification is 7.0 to 7.5

(Gerardi, 2002).

Heterotrophic bacteria degrade organic carbon in order to obtain energy for cellular activity and

carbon for cellular synthesis (growth and reproduction). According to Ekama and Marais (1984),

under anoxic conditions, a theoretical demand of organic biodegradable substrate will be 8.67 mg

COD to reduce 1 mg N-nitrate (Ekama and Wentzel, 2008).

Furthermore, it is mentioned that in practical COD/N ratios required for complete denitrification are 4-

15 g COD/g N, with a minimum ratio of 3.5-4 (Kujawa and Klapwijk, 1999). The Carbon and

Nitrogen (C/N) ratio of domestic wastewaters is often lower than these prescribed values, so that

nitrogen removal is limited by the lack of available organic carbon source (Ryu and Lee, 2009).

Sometimes an external carbon source such as acetate, methanol or ethanol is added in order to achieve

denitrification for ammonia removal. As a result, biological nutrient removal through aerobic

nitrification and anoxic denitrification may increase the operational cost, process complexities and

energy input for aeration.

Dudy Fredy 7

2.2.3 Effect of C/N ratio on nitrogen removal

Addition of carbon substrates increases the COD/NO3-N ratio and may improve denitrification process.

However, denitrification rate may also depend on the types of carbon and quantities of substrate added.

Sodium acetate was known as effective and efficient carbon source, then methanol and glucose. It is

reported that addition of sodium acetate could increase the amount of nitrate reduction even at high

dosage, and improved the rate of nitrogen removal (Tam et al., 1992).

Optimum C/N ratio is the ratio that leads to a maximum conversion of all nitrogen compounds to

nitrogen gas with minimum organic carbon. Theoretical optimal C/N ratio is calculated to be 3.74 for

denitrification system without any competition from other heterotrophs. It also depends on the

characteristic of wastewater being treated. Consequently optimal C/N is not constant and must be

determined experimentally (Chiu and Chung, 2003).

Theoretical optimal C/N ratio may be calculated using the stoichiometric relationship for the

biological denitrification process. The chemical equilibrium equation using acetic acid as carbon

source was suggested as follows (Mateju et al., 1992).

0.819CH3COOH + NO3− → 0.068C5H7NO2 + HCO3

− + 0.301CO2 + 0.902H2O + 0.466N2 (2.4)

2.2.4 SRT (Solids Retention Time)

SRT or sludge age is the most important operational parameter which has been used in the design,

operation and control of an activated sludge systems (Ekama, 2010). The SRT is equal to the mass of

solids in the reactor divided by the mass of solids leaving the system (waste activated sludge solids)

per day.

A successful nitrification processes in both suspended growth or attached biofilm reactors in

wastewater treatment can be determined by SRT. SRT controls microorganism in the system, when the

concentration of microorganisms is high, the SRT is also high (Rittmann and McCarty, 2001). And

when the reactor is in a steady state, SRT is defined as the inverse of the net specific growth rate (μ-kd

).

The nitrification process depends on slow growing autotrophic bacteria. This slow growth rate sets the

minimal value of the SRT in the activated sludge process (Salem et al., 2006). If the SRT was shorter

than the inverse of the specific growth rate (μ-1

), this could cause a washout of nitrifying bacteria. In

most Biological Nitrogen Removal (BNR) processes, a long sludge retention time (8-10 days) is

required due to the slow growth rate of AOB (Ekama and Wentzel, 2008; Lee et al., 2011). On the

other hand, the heterotrophs are known to have higher specific growth rates of around 4 to 13.2 day-1

8 MSc Thesis

than nitrifiers which have specific growth rates of only around 0.62 to 0.92 day-1

respectively (Okabe

et al., 2011).

It is reported that the nitrogen removal efficiency was higher when the SRT increased (Ekama and

Wentzel, 2008). Lee et al. (2008) studied the total nitrogen removal efficiency of an SBR. He

observed that the TN removal efficiency could be obtained up to 66.9% at SRT 16.2 days. However as

the SRT increased, the denitrification rate per mixed liquor suspended solids (MLSS) during the first

anoxic period decreased significantly (Lee et al., 2008). Moreover, another study identified that the

long SRT had bettered the process. It was found that at the SRT between 10.3 to 34.3 days could

lessened the unfavourable effect of low temperatures and stabilized the nitrification process

(Komorowska-Kaufman et al., 2006).

The effect of SRT on flocs sludge characteristics in SBR was studied by Liaou, et al. (2006). It

indicated that floc size was relatively stable and not subject to influence by SRT or organic loading in

SBR. This study also advised that SRT should be larger than the critical SRT (9–12 days) to maintain

a relatively stable microbial community for effective biomass flocculation and separation (Liao et al.,

2006)

In a study of algal-bacterial system in SBR, biomass productivity was generally increased as retention

times decreased (Valigore et al., 2012). And longer SRT enhanced biomass settleability while shorter

Hydraulic Retention Time (HRT) enhanced productivity except when washout occurred (Valigore et

al., 2012)

2.2.5 Sequential Batch Reactor (SBR)

Sequential batch reactor is a modification of activated sludge process. Whereas successfully used to

treat municipal and industrial wastewater (Mahvi, 2008). The difference is that the SBR performs

equalization, biological treatment, and secondary clarification in a single tank/reactor using a timed

control sequence.

In SBR, treatment of wastewater is carried out in various consecutive phases namely filling, reaction,

settling, and decant (as shown in Figure 2.1). The removal of nitrogen can be achieved by alternating

aerobic and anoxic periods during the reaction (Rodríguez, Pino, et al., 2011). The duration of each

cycle and the number of stages of operation depends on the type of wastewater to be treated

(Rodríguez, Ramírez, et al., 2011). The advantages of operation in SBR are single-tank configuration,

small foot print, easily expandable, simple operation and low capital costs (Mahvi, 2008).

Dudy Fredy 9

Figure 2.1:Typical cycles in Sequential Batch Reactor

Source: (Mahvi, 2008)

Many studies have been done with the different operational sequences as shown in Table 2.3, and

generally the objectives were to optimize nitrogen removal.

Table 2.3: Different operational sequences of SBR operation

Operational mode Operational overview Reference

Intermittenly aerated

SBR to treat high

phenol concentration

(Singh and

Srivastava,

2011)

Automatically

controlled SBR to

enhance nitrogen

removal, step-feed

strategy, without

external carbon

source

(Puig et al.,

2005)

Automatically

controlled SBR to

enhance nitrogen

removal, step-feed

strategy, with external

carbon source

Guo et al.

(2007)

10 MSc Thesis

Operational mode Operational overview Reference

Automaticallay

controlled (real time)

SBR to remove

nitrogen via nitrite,

external carbon

source

(Wu et al.,

2011)

A pilot scale SBR to

treat high amount of

organic matter and a

high amount of

ammonium

(Rodríguez,

Ramírez, et al.,

2011)

Source :(Windraswara, 2013)

Irvine and Bush (1979) reported that SBR is an effective biological treatment method for removing

organic matter and nutrients. It could be done by distributing the influent injection and aeration

periods variably and appropriately. In particular, a higher efficiency of denitrification can be achieved

in the SBR method by varying the proportional distribution of the durations of the anoxic and aerobic

periods during one-cycle operation. Lee et al., (2007) indicated that increasing the duration of the

anoxic (II) period, which is conducive to denitrification, increases the efficiency of nitrogen removal

by denitrification.

2.3 Algal-bacterial consortium for wastewater treatment

The application of algal–bacterial biomass for wastewater treatment now is becoming more interesting.

It presents lower energy consumption via photosynthetic aeration (Muñoz et al., 2004; Safonova et al.,

2004; Muñoz and Guieysse, 2006), and offers the potential use as an alternative energy source (biofuel

or biogas) from its biomass. Furthermore the algal based system may also contribute to CO2 mitigation

(Muñoz and Guieysse, 2006; Subashchandrabose et al., 2011).

The interactions of microalgae and ordinary heterotrophic (OHOs) bacteria in wastewater treatment

process can be a symbiotic relationship. Microalgae provide the necessary O2 for heterotropic bacteria

to biodegrade organic pollutants, and the CO2 released from bacterial respiration is used for

photosynthesis. As autotrophic nitrifiers and hetereotrophic denitrifers are also present in the system,

the interactions between microorganisms become more complex, as shown in Figure 2.2.

Dudy Fredy 11

Figure 2.2:Interaction between algae and bacteria in wastewater treatment process

Moreover, microalgae and bacteria do not limit their interactions to a simple CO2-O2 interchange.

Microalgae may increase bacterial activity by releasing extracellular (Wolfaardt et al., 1994). While,

De-Bashan et al. (2002) reported that the presence of growth promoting bacteria Azospirillum

brasilense enhanced ammonium and phosphorous removal by C. vulgaris (de-Bashan et al., 2002).

However microalgae may also have an unfavorable effect on bacterial activity by increasing the pH,

the DO (Dissolved Oxygen) concentration, or the temperature of the cultivation broth, or by excreting

inhibitory metabolites (Oswald, 2003; Schumacher et al., 2003). Or even the other way around

bacteria may inhibit microalgae by producing algicidal extracellular metabolites (Fukami et al., 1997).

Nevertheless, treatment of wastewater using the interaction of algal-bacterial was developed. W.J.

Oswald in the year 1950s introduced such kind technology for sewage treatment (Oswald and Gotaas,

1957). Since then, many improvements have led to the use of algal-bacterial consortium in facultative

ponds, high rate algal ponds (HRAP), and closed photo-bioreactors (Babu et al., 2010; Park et al.,

2011; Subashchandrabose et al., 2011; Craggs et al., 2012). Table 2.4 summarizes various studies that

reported results on nutrient removal by algal-bacterial consortium to treat wastewater.

ALGAE

OHOs

NITRIFIERS

DENITRIFIERS

LIGHT

Wastewater BIOMASS

CO2

N

CO2

N

NO3

Organics

O2

Reclaimed Water

CO2

12

M

Sc T

hesis

Table 2.4: Algal-bacterial consortium for nutrient removal from wastewater

Cyanobacterium/microalga Bacterium Source of wastewater Nutrients

Removal

Efficiency

(%)

Initial conc

(mg/l) Reactor used

Spirulina platensis Sulfate-reducing bacteria Tannery effluent Sulfate 80 2000 High rate algal pond (HRAP)

Chlorella vulgaris Azospirillum brasilense Synthetic wastewater Ammonia 91 21 Chemostat

C. vulgaris Wastewater bacteria Pretreated sewage DOC 93 230 Photobioreactor pilot-scale

nitrogen 15 78

C. vulgaris Alcaligenes sp. Coke factory wastewater NH4+ 45 500 Continuous photobioreactor

with sludge recirculation phenol 100 325

C. vulgaris A. brasilense Synthetic wastewater Phosphorous 31 50 Inverted conical glass bioreactor

nitrogen 22 50

Chlorella sorokiniana Mixed bacterial culture

from an AS

Swine wastewater Phosphorous 86 15 Tubular biofilm photobioreactor

nitrogen 99 180

C. sorokiniana Activated sludge bacteria Pretreated piggery wastewater TOC 86 645 Glass bottle

nitrogen 87 373

C. sorokiniana

Activated sludge

consortium Pretreated piggery slurry TOC 9 to 61 1247

Tubular biofilm photo

bioreactor

nitrogen 94-100 656

Phosphorous 70-90 117

C. sorokiniana Activated sludge bacteria Piggery wastewater TOC 47 550 Jacketed glass tank

photobioreactor phosphorous 54 19

NH4+ 21 350

Euglena viridis Activated sludge bacteria Piggery wastewater TOC 51 450 Jacketed glass tank

photobioreactor phosphorous 53 19

NH4+ 34 320

Abbreviations: DOC=dissolved oxygen concentration; TOC=total organic carbon

Source: (Subashchandrabose et al., 2011)

Dudy Fredy 13

As background information for using microalgae for wastewater treatment, the following sub-sections

discuss characteristics and limiting factors for algal growth.

2.3.1 Microalgae

Micro-algae may be unicellular or multicellular, and can be found in all water bodies (e,g. fresh-water,

sea-water and hypersaline lakes). They are also found in soils, on plants (terrestrial and aquatic) and

form symbiotic associations with a very wide range of plants and animals.

Microalgae may incorporate in many types of metabolisms (i.e. autotrophic, heterotrophic,

mixotrophic, photoheterotrophic) and able to have a metabolic shift as a response of changes in the

environment conditions (Mata et al., 2010). Two major types of its metabolisms are photoautotrophic

and heterotrophic. Photoautotrophs obtain all the elements they need from inorganic compounds and

the energy for their metabolism from light. Heterotrophs obtain their material and energy needs from

organic compounds synthesized by other organisms.

2.3.2 Microalgae culture system

Currently, various types of bioreactors have been used for culturing algae. They have to provide

suitable conditions of temperature, pH, mixing, and substrate concentration for efficient cellular

metabolism. The highly controlled culture systems are known as photo-bioreactors (PBR). They can

be open or closed culture systems.

Open photo-bioreactors (ponds) are less expensive, more durable, and higher productivity than large

closed reactors. However, ponds are more sensitive to weather conditions, evaporation, lighting and

water temperature. They also are more prone to contaminations from other microalgae or bacteria, and

need more land area (Mata et al., 2010). The most widely used open systems are waste stabilization

ponds (WSP) and high rate algal ponds (HRAP).

Closed photo-bioreactors have several advantages over open photo-bioreactors. They provide better

control over growth parameters (temperature, CO2, pH, mixing, and O2), culture conditions, reduces

CO2 losses, prevent evaporation, allows higher microalgae densities, offer a more safe, higher

volumetric productivities, and protected environment (Mata et al., 2010). They also are flexible for

being optimized for appropriate biological and physiological characteristics of the cultivated algal

species.

14 MSc Thesis

2.3.3 Influence of environmental parameters on algal growth

The growth rate of microalgae is influenced by physical (e.g. light and temperature), chemical (e.g.

availability of nutrients, carbon dioxide, pH), biological factors (e.g. competition between species,

grazing by animals, virus infections) and operational factors (e.g. bioreactor design, mixing and

dilution rate) (Larsdotter, 2006). Some of the most important parameters are described in below sub-

sections.

2.3.3.1 Light

Microalgae are phototrophs, they obtain energy from light. The light energy is converted to chemical

energy in the photosynthesis, but large parts are lost as heat. Oswald (1988) reported that in outdoor

ponds, more than 90% of the total incident solar energy is converted into heat and less than 10% into

chemical energy. Fontes (1987) reported a conversion efficiency of sunlight energy into chemical

energy of only 2% (Larsdotter, 2006).

The relationship between light intensity and photoautotrophic cell growth or various cell activities is

rather complex. Algal activity increases with light intensity up to 200–400 µE/m2.s (Muñoz and

Guieysse, 2006), while insufficient light lowers the growth rate. Anexcess light may lead to

photoinhibiton (Hsieh and Wu, 2009). Photoinhibition has therefore been observed at noon of a sunny

day when irradiance can reach more than 4000 µE/m2.s (Fuentes et al., 1999; Carvalho et al., 2011). A

typical relation between light intensity and photoautotrophic growth is shown in Figure 2.3.

Figure 2.3: Relation between light intensity on photoautotrophic growth of photosynthetic cells

Source: (Ogbonna and Tanaka, 2000)

Dudy Fredy 15

Absence of light or low light intensity can also lower photosynthesis efficiency, and it will leads to

anaerobic conditions in the reactor (Muñoz and Guieysse, 2006). Some algae are able to grow in the

dark using simple organic compounds as energy and carbon source (heterotrophic) (Perez-Garcia et al.,

2010), and their cells metabolize their components to obtain maintenance energy, thus leading to a

decrease in cell weight (Ogbonna and Tanaka, 2000).

2.3.3.2 Temperature

The growth rate of algae increases with the increased of temperature. As cited by Park et al., (2011),

Soeder et al., (1985) observed that the optimal temperature for maximum algal growth rate (sufficient

nutrient and light conditions) varies between algal species. The optimal temperature for many algae is

between 28 and 35°C. While Harris (1978) reported that optimal temperature also varies when nutrient

or light conditions are limited. The algae growth often declines when encounter to a sudden

temperature change (Park et al., 2011).

Temperature is also well connected with light intensity. As cited in (Muñoz and Guieysse, 2006),

Abeliovich (1986) reported that excessive temperature can happen at high light intensities and high

biomass concentrations, from the fact that algae convert a large fraction of the sunlight into heat.

2.3.3.3 Nutrients

Algae nutritional requirements can be classified into macronutrients (i.e. those required at g/L

concentrations) and micronutrients (i.e. those required at mg/L or μg/L concentrations). The primary

essential elements for the growth of algae are carbon, nitrogen and phosphorus. Algae also need some

trace elements including calcium, iron, silica, magnesium, manganese, potassium, copper, sulfur,

cobalt and zinc which are also essential but rarely affect its growth in wastewater treatment

(Christenson and Sims, 2011).

Ammonium and nitrate are the most commonly inorganic nitrogen sources in algal media. Some of the

prokaryotic algae, the blue-green algae (cyanobacteria), can also fix atmospheric N2. The pH of the

medium may decreased when ammonium is used as the only nitrogen source. In dense cultures and

high temperatures, the lower pH may result into a rapid decline in growth or even death of the culture.

It is also reported that high concentrations (more than one mM) ammonium may inhibit growth,

especially at high temperatures (Borowitzka, n.d.).

If both ammonium and nitrate are supplied the cultures generally do not take up the nitrate until the

ammonium has been used up. This is because ammonium is the end product of nitrate reduction,

16 MSc Thesis

therefore it may cause feedback-inhibition and repression of the nitrate uptake and reduction system

(Borowitzka, n.d.)

Phosphorous is another major nutrient for algae, as an inorganic form of phosphate (H2PO4- and

HPO42-

). The take up is started by hydrolysis with the action of phosphoesterase or phosphatase

enzymes The amount of inorganic phosphorus required by algae varies among algae species. The

tolerance amount is normally around 50 μg/L – 20 mg/L (Becker, 1994).

2.3.3.4 pH

pH may have a great effect on the microalgae growth rate and species composition. For example, it

was found that the optimal productivity of Anabaena variabilis (cyanobacterium) was achieved at pH

8.2 – 8.4 and decreased slightly at pH 7.4 – 7.8, while at pH 9.7 – 9.9 the cell could not survive

(Larsdotter, 2006).

Photosynthetic assimilation (CO2 uptake by algae) may also increase pH in the medium. The increase

of pH can reach 11 or more if CO2 is limiting and bicarbonate is used as a carbon source (Larsdotter,

2006).

Nitrogen assimilation by the algae also affects pH. Assimilation of nitrate ions tend to raise the pH,

but if ammonia is used as nitrogen source, the pH of the medium may decrease to as low as 3 (Becker,

1994). At high pH (above 9) free ammonia will begin to dominate over ammonium. High

concentrations of ammonia may inhibit algal growth, and this toxicity is intensified at higher

temperatures which may freely diffuse over membranes into the cells (Larsdotter, 2006; Markou and

Georgakakis, 2011).

2.3.3.5 Dissolved Oxygen (DO)

The process of photosynthesis is the main source of DO in outdoor ponds. The amount of DO

concentration depends on the growth rate of algae, light intensity and temperature; it reaches a

maximum value during noon and then decreases as light and temperature decrease. A study using

P.carterae showed that the maximum concentration of O2 during the day was directly related to light

irradiance. The highest photosynthetic rate was obtained in high light (1,900 μmol photons/m2.s) and

at low oxygen concentration (6–10 mg O2/L). However, high concentrations of oxygen (26–32 mg

O2/L) at both low and high irradiances, strongly increase the degree of inhibition of photosynthesis

(Moheimani and Borowitzka, 2007).

Dudy Fredy 17

2.3.3.6 Carbon

Microalgae obtain energy through photosynthesis. As indicated in the photosynthetic equation 2.4,

solar energy is converted into chemical energy and this energy is then used to assimilate inorganic

carbon (CO2) to produce sugars and oxygen as a by-product.

6 H2O + 6 CO2 + light 6 C6H12O6 + 6 O2 (2.4)

Inorganic (CO2 and HCO3-) and organic carbon are utilized by algae in the ponds as well as in algal

culture systems. The organic carbon sources can be assimilated either chemo- or photo-

heterotrophically (Larsdotter, 2006). In the first case, the organic substrate is used both as the source

of energy (through respiration) and as carbon source, while in the second case, light is the energy

source. In several algal species e.g. Chlorella and Scenedesmus, the mode of carbon nutrition can be

shifted from autotrophy to heterotrophy when the carbon source is changed (Becker, 1994).

The amount of CO2 dissolved in water varies with pH, where an addition of CO2 results in a decreased

pH. At higher pH values, for example at pH greater than 9, most of the inorganic carbon is in form of

carbonate (CO3 2–

) which cannot be assimilated by the algae. The decreased availability of CO2 may

act as a limiting factor on the algal growth, however his effect is not often evident. Aeration is one of

few ways in providing carbon dioxide for algae growth (Larsdotter, 2006).

A study observed in Chlorella sp. cultivation in photo-bioreactors showed that 2% aeration of CO2

increased the total biomass productivity (Chiu et al., 2008). The CO2 demand for microalgae can also

be calculated based on its stoichiometric requirement because 1 g biomass requires 1.85 g CO2. For

example, if the growth rate of biomass is 1 g/(g.d) at biomass concentration 1 g/L, then the carbon

transfer rate needed would be 1.85 g/(L.d) (Posten, 2009).

2.4 Effect of dark periode on nitrogen removal

Irradiance or light is one of the major factors, which control microalgae productivity. Absence of light

can cause a reduction of photosynthesis activity, which normally leads to the occurrence of anaerobic

conditions in the reactor. This condition is assumed to be favored for denitrification process. In

activated sludge system, denitrification has been observed to begin when DO concentration in the

flocs drops below 0.6 mg/L (Schramm et al., 1999).

On the other hand, there are a number of microalgae that can grow mixotrophically. Mixotrophic

growth regime is a variant of the heterotrophic growth regime, where CO2 and organic carbon are

18 MSc Thesis

simultaneously assimilated and both respiratory and photosynthetic metabolism operate concurrently

(Perez-Garcia et al., 2011). In the absence of light, there may be a competition between microalgae

and bacteria for an organic carbon source, which may result in a negative effect on the denitrification

process.

Dudy Fredy 19

3 Objectives

3.1 General objective

To optimize nitrogen removal in a Sequential Batch Photo-bioreactor fed with artificial UASB effluent

by variations in operational conditions.

3.2 Specific objectives

1. To study the effects of different operational sequences in an SBR cycle on the nitrification and

denitrification capacity of microalgae and bacteria consortium in a Sequential Batch Photo-

bioreactor.

2. To study the effect of SRT on the nitrification and denitrification capacity of microalgae and

bacteria consortium in a Sequential Batch Photo-bioreactor.

20 MSc Thesis

Dudy Fredy 21

4 Material and Methods

The research was carried out in UNESCO-IHE laboratory in Delft, the Netherlands and it was a

continuation of previous MSc thesis research by UNESCO-IHE student Rudatin Windaswara

(Windraswara, 2013). The photo-bioreactors was operated in the SBR mode, controlled by a

bioconsole system, and connected to a data logger for DO concentration and pH measurement. The

work included operating the system, preparing artifical wastewater, sampling and analyzing nitrogen

compounds, chlorophyll-a, VSS/TSS, COD, and light absorption. Nitrogen conversion rates are

determined and nitrogen mass balances are developed.

4.1 Culture medium

The artificial UASB’s effluent was used as culture medium and done by modifying a microalgae

medium BG-11 (Stanier et al., 1971), similar to the medium used on previous study by Windraswara

(2013). The sole nitrogen source was ammonium sulphate (NH4)2SO4, with the concentration of

nitrogen about 23 mg N-NH4+/L. Sodium acetate was used as the Carbon source with the concentration

about 200 mg COD/L. As trace elements source, the Ogawa solution was applied to the medium for 2

mL.To avoid precipitation, MgSO4 and CaCl2.2H20, and COD source were prepared in different

reservoir flasks. The detailed compounds are shown in Table 4.1.

Table 4.1: Modified BG-11 medium for microalgae and bacteria culture

No Compounds Concentration (g/l) 1 C6H8O7.H20 0,0066 2 (NH4)2SO4 0,1094 3 K2HPO4 0,1302 4 MgSO4 0,0750 5 CaCl2.2H2O 0,0360 6 NaHCO3 0,4200 7 FeSO4.7H2O 0,0034 8 Na2CO3 0,0200 9 Na2SiO3.5H2O 0,0448

10 EDTA.Na2 0,0010 11 COD source (CH3COONa) 0,4240 Ogawa 2 solution = 2 ml Stock solution (Ogawa 2) 1 H3BO4 3,5875 2 MnCl2.4H2O 1,8100 3 ZnSO4.7H2O 0,2200 4 CuSO4.5H2O 0,0800 5 (NH4)6Mo7O24.4H2O 1,2879 6 CoCl2.6H2O 0,0340 7 NiCl.6H2O 0,0430 8 KI 0,1800 9 EDTA.Na2 0,0800

22 MSc Thesis

4.2 Microalgae-bacteria consortium

Biomass culture was consisted of algae species Scenedesmus quadricauda, Chlorella sp, Anabaena

variabilis, Chlorococcus sp, and Spirulina sp, enriched with wild algae species from a canal in Delft,

the Netherlands. Nitrifying and denitrifying bacteria population were taken from fresh sludge of

Harnaschpolder Wastewater Treatment Plant (WWTP), Delft, the Netherlands.

4.3 Reactor set up and experimental design

The experiment was conducted in 1-L cylindrical jacketed and transparent glass reactor. The

temperature and pH during React phase was kept at 28oC and 7.5, respectively. Four sets of standing

lamps (Phillips, the Netherlands) were installed at four sides of the reactor (average light intensity on

the surface of reactor’s wall of 25.9 µmol/m2s). Mixing was maintained at 200 rpm during Fill and

React phases, and DO concentration and pH were measured and recorded using the DO probe from

Bio-console Applikon, the Netherlands. The schematic diagram of reactor set-up is shown in Figure

4.1.

Figure 4.1: Schematic diagram of reactor set up

Source: (Karya et al., 2013)

Dudy Fredy 23

The reactor was operated as an open system of SBR, with two cycles per day (1 cycle in 12 hours).

The experimental variations were done in four periods. The first two periods were done in different

sequential operation and the other three periods were done in different SRT as described in Table 4.2.

Table 4.2: Experimental variations

Period Sequential operation in one cycle of SBR mode SRT

(days)

1

48

2

52

3

26

4

17

The more detailed operational setting of one cycle for each period is described in Table 3.3 and Table

3.4. Withdrawal of the effluent was 50% of the total reactor volume. The duration of each SBR phases

were controlled using a set of automatic controllers from Bio-Console Applikon, Holland.

Table 4.3: SBR operational setting at period 1

No Phase Time

(min) Set

pH Light Stirrer Remark

0 initialization 0 NA on

on

1 start 2 7.5

on

2 filling 1 8 7.5 400 ml of substrate

3 aerobic 1 440 7.5

4 pre anoxic 15 7.5

off

5 filling 2 2 7.5 100 ml of COD source

6 anoxic 73 7.5

7 aerobic 2 60 7.5

on

8 settling 105 NA

off

9 withdrawing 10 NA

10 iddle 5 NA

Total time 720

Influent

WithdrawFilling Aerobic 1 Anoxic Aerobic 2 Settling

Carbon source

Withdraw

Filling 1

Aerobic 1Anoxic 1 Aerobic 2 Settling

Influent and

Carbon source

Anoxic 2

Influent and

Carbon source

Filling 2

24 MSc Thesis

Table 4.4: SBR operational setting at period 2, 3 and 4

No Phase Time

(min) Set

pH Light Stirrer Remark

0 initialization 0 NA on

on

1 start 2 7.5

off

2 pre-anoxic 1 13 7.5

3 filling 1 4 7.5 200ml of substrate + 50

mL of COD

4 anoxic 1 26 7.5

5 aerobic 1 255 7.5 on

6 pre-anoxic 2 15 7,5

off

7 filling 2 4 7.5 200ml of substrate + 50

mL of COD

8 anoxic 2 26 7.5

9 aerobic 2 255 7.5

on

10 settling 115 NA

off

11 withdrawing 4 NA

12 iddle 1 NA

Total time 720

4.4 Sampling

The samples for daily influent nitrogenous and COD concentrations were taken from the influent

reservoirs at the beginning of the cycle. While the effluent samples were collected during the withdraw

phase at the end of the cycle. For light absorption, TSS/VSS and chlorophyll-a analysis, the samples

were collected after second filling phase completed. For all nitrogenous analysis, samples are filtered

directly after the collection using 0.45 μm filter, and then stored in the refrigerator (4oC) for the

analysis.

4.5 Analytical Methods

4.5.1 Ammonium nitrogen

Based on NEN 6472 method, ammonium nitrogen concentration was measured using the

spectrophotometer. Samples were filtered over a glass fiber (GF/C), pipetted into a 50 mL flask and

reacted with sodium salicylate reagents and dichloroisocyanurate reagents. A series of standards of

NH4Cl solutions were used from standard solution with known concentrations to develop a calibration

curve. By using the spectrophotometer, the absorbance of each standard sample and the samples were

measured at the wavelength of 655 against water with 1 cm cells between 1 to 3 hours. The results of

Dudy Fredy 25

these standards were plotted against their known concentrations to determine the mathematical

expression which further were used to determine the concentration of samples from the experiment.

4.5.2 Nitrite nitrogen

Nitrite nitrogen concentration was determined according to Standard Methods for examination of

water and wastewater from American Public Health Association (APHA, 1995). The analysis was

using the colorimetric procedure which employs two organic reagents, namely sulfanilamide and N-(1

Naphtyl)-ethylenediamine dihydrocloride. An amino group from sulfanilamide reacts with nitrite ion

as nitrous acid which resulted in a pinkish-red azo dye. A series of NO2- standards were used to

develop a calibration curve in each analysis of the samples. Diluted samples were mixed with 2 mL of

mixed reagents in 50 mL flasks after which the photometric measurement was conducted at the

wavelength of 543 nm against water with 1 cm cell (between 10 minutes to 2 hours).

4.5.3 Nitrate nitrogen

Ion chromatography method using Dionex ICS-1000 is used to determine nitrate nitrogen

concentration. Samples were filtered through 0.45 μm filters immediately after collection to prevent

the bacteria in the sample that may change the ionic concentrations. Dilution of samples by using de-

ionized water was needed to avoid high concentrations of nitrite nitrogen not greater than 10 mg/L.

For analysis, 5 mL of samples was placed in high density polyethylene containers and washed

thoroughly using de-ionized water.

4.5.4 Chlorophyll-a

Chlorophyll-a concentration was determined according to Dutch Standard NEN 6520. Samples were

filtered by using GF6 filter (0.45 μm porosity) and transferred to Schott GL 18 COD tubes. The

chlorophyll-a was extracted using 80% (v/v) ethanol. To achieve complete pigment extraction, brief

heating for about 5 minutes in a water bath at temperature of 75oC was conducted. To promote better

extraction, furthermore the tubes were shaken several times and cooled to room temperature. After

centrifugation, the supernatant was analyzed in spectrophotometer at a wavelength of 665nm against

80% ethanol. For turbidity correction, measurement at a wavelength of 750 was also carried out. After

obtaining the absorbance reading, 2 drops of 0.4 M HCl were added to each sample and (5 to 30)

minutes later the samples were re-measured at the same wavelengths. The following equation was

used to calculate the chlorophyll-a concentration:

Chl-a (μg/L) = 296 * V1 * En-Ea/(Vo * p) (4.1)

26 MSc Thesis

Where En=Ex-Eo is the corrected absorbance of the non-acidified extract, Ea=Exa-Eoa is the

corrected absorbance of the acidified extract, 296 is a correction factor on the specific absorption

coefficient of Chl-a, V1 is the volume of 80% ethanol in mL, Vo is the sample volume in L and p is

the cell thickness in mm.

4.5.5 Total suspended solid (TSS) and volatile suspended solid (VSS)

The determination of TSS and VSS involved of drying the samples at 105oC in an oven and

combustion at 520oC in a muffle furnace according to APHA (1995). Well mixed samples were

filtered using weighed GF/C filters which had been pre-heated for 2 hours at 520oC and stored in a

desiccator. Filtered sludge solids were placed in aluminum cups and left in the oven at 105oC for 2

hours then weighted. The dry weight (TSS) is calculated by substracting the dry weight (sample+filter)

from a clean weight (filter). Samples were put back in the cups, combusted at 520oC for 3 hours and

cooled in a desiccator before weighing. The difference in weight before and after combustion

represented the VSS (g/L).

4.5.6 Biomass Light Absorption

The well mixed of liquor sample from the reactor was consisted of biomass with certain chlorophyll-a

concentration. Its light absorption was determined by using a spectrophotometer in a 1 cm cuvette.

The initial light intensity was measured with an LI-COR LI-1400 quantum sensor. Repeating the

procedure with a cuvette blanked with medium, the biomass light aborption coefficient could be

determined by rearranging equation Beer-Lambert’s Law as:

ka =

(4.2)

where I is a light intensity at distance z, Io is the incident light intensity at the reactor, B is the biomass

concentration, and ka is the biomass light absorption coefficient.

4.6 Calculation

4.6.1 Solid Retention Time (SRT)

The actual SRT was determined based on formula in a study done by (Valigore et al., 2012). This

formula also considers the biomass concentration in wastage during discharge. The formula is shown

below:

Dudy Fredy 27

(4.3)

Where: Xr biomass concentration in the reactor (mg/L)

Xs biomass supernatant (mg/L)

Qw waste dischare flowrate (L/day)

Qs supernatant discharge flowrate (L/day)

4.6.2 Nitrogen balance

Nitrogen balance in the reactor was obtained from the overall nitrification equation by using the

stoichiometric link developed by Liu and Wang in their study (Liu and Wang, 2012). The equation

provides a more accurate stoichiometric link between nitrifier yield, ammonia consumption, and

oxygen uptake for both steps of the nitrification process. And also considers the amounts of ammonia

incorporated into the cells of ammonia oxidizers and nitrite oxidizers.The complete nitrification

equation is shown below:

NH4+ + 0.0298NH4

+ + 1.851O2 + 0.1192CO2 + 0.0298HCO3

-

0.0298C5H7O2N + NO3- + 0.9702 H2O (4.4)

Initial ammonium concentration

Initial ammonium concentration was obtained from the measurement of influent concentration in the

reservoir and multiplied it by dilution factor.

Ammonium uptake by nitrifiers (AOB and NOB)

Equation 4.4 shows that when 1 unit of ammonia is oxidized into nitrite then into nitrate, 0.0298 unit

of ammonia-nitrogen would be incorporated into nitrifiers.

Nitrified ammonium

The highest nitrate concentration in react phase (sum of aerobic 1 and aerobic 2 for period 2 to 4) has a

correlation with the ammonium nitrogen that had been oxidized (nitrified ammonium). The nitrified

ammonium was calculated from the Equation 4.4 which shows that 1 mg N-NO3- is formed from (1+

0.0298) mg N-NH4+.

Ammonium uptake by algae

Since the initial ammonium nitrogen concentration is higher than the amount that had been nitrified, it

means the difference was assumed as ammonium that had been uptaken by algae. And it can be

assumed that there was no NO3- uptake by algae.

28 MSc Thesis

4.6.3 Biomass composition

There are at least four different organisms that contributed to the total biomass in the reactor of algae-

bacteria consortia. Those four microorganisms namely are autotrophic nitrifiers, heterotrophic

denitrifiers, ordinary heterotrophic organisms (OHO) and algae itself. The calculations of biomass

production are described below:

Nitrifiers biomass production

Nitrifiers biomass production was obtained from Equation 4.4, where 1 mg N-NO3- produced 0.24 mg

of biomass (C5H7O2N).

Denitrifiers biomass production

The denitrifiers organisms biomass was calculated stoichiometrically based on nitrate formed during

the reaction phase. From the denitrification equation (Mateju et al., 1992) which used acetate as a

carbon source, 0.55 mg of biomass (C5H7O2N) is produced from 1 mg N-NO3- , as shown below:

0.819CH3COOH + NO3− → 0.068C5H7NO2 + HCO3

− + 0.301CO2 + 0.902H2O + 0.466N2 (4.5)

OHO biomass production

Ordinary heterotrophs biomass was obtained through the stoichiometry of acetate oxidation equation.

The biological reaction for acetate oxidation is adapted from Metcalf and Eddy (2003) (Henze, 2008),

by combining three different half reactions of cell synthesis, electron acceptor and electron donor. The

overall balanced equation is shown below:

0.125CH3COO- + 0.0295NH4

+ + 0.103O2→ 0.0295C5H7NO2 + 0.095HCO3

− + 0.007CO2 + 0.955H2O

(4.6)

Based on the measurement of influent and effluent of COD, and from COD requirement for

denitrification through Equation 4.5, the COD that had been used for acetate oxidation can be known.

Furthermore the OHO biomass production can be calculated.

Algae biomass production

The algae biomass production was calculated from the photosynthesis reaction (Mara, 2003) below:

106CO2 + 236H20 + 16NH4+ + HPO4

2- C106H181O45N16P + 118O2 + 171H2O +14H

+ (4.7)

Dudy Fredy 29

The biomass production was calculated stoichiometrically based on the amount of ammonium uptake

by algae as described in section 4.6.2.

4.6.4 Oxygen production by algae

The estimation of oxygen production by algae was developed from the oxygen mass balances in one

cycle of SBR operation. The oxygen concentration profile of one cycle operation was divided into

three phase, as illustrated in Figure 4.2.

Figure 4.2: DO concentrations profile partition in one cycle of SBR operation

Phase I

It started from the beginning of the cycle until the DO concentration reach near zero mg/L. It assumed

that the respiration by algae was negligible. In this phase, the oxygen mass balance is shown below:

(Cs-C) + ralg – rnit – rhet (4.8)

Phase II

It occurred when the DO concentration reach near zero mg/L. In this phase, the oxygen mass balance

is developed with assumed that the respiration by algae was negligible, and the organic material

oxidation had completely finished.

(Cs-C) + ralg – rnit = 0 (4.9)

0

2

4

6

8

10

12

14

16

18

20

0 1 2 3 4 5 6 7 8 9 10 11 12

DO

co

nce

ntr

atio

n (

mg/

L)

Time (h) phase I

phase II phase III

slope of phase III

slope of phase I

30 MSc Thesis

Phase III

It occurred when the DO concentration rose until the second filling time. In this phase, the oxygen

mass balance assumed that the respiration by algae was negligible, and the organic material oxidation

and nitrification has completed.

(Cs-C) + ralg (4.10)

Where:

C DO concentration (mg/L)

Cs Saturation DO concentration (mg/L)

KLa oxygen mass transfer coefficient (1/hour)

ralg rate of oxygen generation by algae through photosynthesis (mg/hour)

rnit rate of oxygen consumption for nitrification (mg/hour)

rhet rate of oxygen consumption for organic matter oxidation (mg/hour)

The calculation procedures are described as follows:

a. Equation 4.10 can be derived as follow:

ralg

where r = KLa*Cs + ralg

k = KLa

=> -(1/k)·ln(r - k·C) = t + a (a is constant of integration)

r - k·C = e^(-k·(t+a)), after using initial condition to evaluate a

C(t) = r/k - (r/k - C₀)·e^(-k·t)

C(t) = r/k - (1 - e^(-k·t)) + Co* e^(-k·t)

C(t) = A - (1 - e^(-B·t)) + Co* e^(-B·t) (4.11)

b. Determine A and B, through curve fitting of equation 4.11 with data of experiments. The

curve fitting was done with online software in http://zunzun.com/

c. Determine the ralg from the value of A and B

d. Solution of equation 4.9 and 4.8 were done in similar way.

Dudy Fredy 31

4.7 Statistical analysis

Statistical analyses were performed using free software R Studio, version 0.97.551. Normality of the

data and homogeneity of variances were determined prior to any statistical treatments with a Shapiro-

Wilk’s test and Q-Q plots, respectively. And the normal distribution and homogeneity of variances

were not observed. The significant differences between mean values were analysed by Mann–Whitney

U-test, for period 1 and 2. The analysis was followed by Kruskal–Wallis H-test and a Tukey’s post

hoc test to see the significant differences between mean values of period 2, 3, and 4. Moreover, all

means were given as mean with standard deviation (number of measurements).

4.8 Batch experiment

To have a better understanding on nitrate uptake by algae in the reactor, batch experiment was

conducted.The same biomass culture and medium was used for the batch experiment. In the medium

the only Nitrogen source was supplied from NaNO3. Of each flask, 50 mL of the above culture was

grown in 200 mL of modified BG-11 medium in duplicate to gain three treatments of 0, 9, and 18

mg/L N-NO3-. The flasks were placed on a shaker with 140 rpm at room temperature. The lamp was

positioned above the shaker with 60 µmol m-2

s-1

photon. To avoid denitrification process occurred,

the light was always turned on for a period 24 hours per day. N-NO3- concentrations were measured

daily. After the entire N-NO3- was diminished in the flasks, refill of N-NO3

- was done to observe the

NO3- uptake performance.

32 MSc Thesis

Dudy Fredy 33

5 Results

5.1 Nitrogen removal

5.1.1 Daily nitrogenous concentration

The daily nitrogenous concentration profile for all periods is shown in Figure 5.1. Nitrification was

achieved completely without external aeration. A high concentration of nitrate in the effluent was

observed at the end of period 1 and within period 2. It was caused by insufficient COD source for

denitrification process, due to high algae growth in period 1 and COD supply failure in period 2. The

explanation of why COD could be the cause is discussed in section 6.1.

Figure 5.1: Daily nitrogenous concentration

The ammonium and total nitrogen removal efficiency is shown in Table 5.1. In general, the removal

efficiency for ammonium was 100% and for total nitrogen was 90%. As mentioned before due to

insufficient COD in the reactor, the nitrogen removal efficiency in period 2 was low 68%. However

there was no significant difference of nitrogen removal efficiency between all periods (P>0.05).

Table 5.1: Nitrogen removal efficiency

0.0

5.0

10.0

15.0

20.0

25.0

30.0

0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170

Nit

roge

no

us

con

cen

trat

ion

(m

g/L)

Days

Inf N-NH4 Eff N-NH4 Eff N-NO2 Eff N-NO3

Period Ammonium Removal

Efficiency (%) Total Nitrogen Removal

Efficiency (%)

Avg Std Dev Avg Std Dev 1 100 0.1 92.7 21.2

2 100 0.0 68.4 24.7

3 99.9 0.4 89.5 1.5

4 99.6 0.9 90.7 6.1

Period 1 Period 2 Period 3 Period 4

34 MSc Thesis

5.1.2 Ammonium conversion rate

A typical nitrogeneous concentration profile in one cycle of operation in the SBR system for period 2

is shown in Figure 5.2. From the linear part of ammonium conversion curves, conversion rate constant

in two different aerobic phases of the SBR operation were calculated. These rates were calculated

from the slope (ka and kb) where ka denotes ammonium conversion rate constant in first aerobic phase

and kb denotes ammonium conversion rate constant in the second aerobic phase.

Figure 5.2: Ammonium conversion in one cycle of SBR operation in period 2 day 89

The ammonium conversion rates for all periods are summarized in Table 5.2. The statistical tests were

done to see the differences of specific ammonium rate between each period. A significant difference

between the specific ammonium conversion rate of period 1 and period 2 could be demonstrated

(W=24, P<0.05).

Table 5.2: Summary of ammonium conversion rate in different period

Period Day Rate of ammonium conversion

(mgN-NH4/L.h) VSS

Rate of ammonium conversion (mgN-NH4/gVSS.h)

ka kb average k Std Dev g/L ka kb average k Std Dev

1

4 -2.5

-2.8 0.4

1.1 -2.3

-2.4 0.2 9 -2.7 1.1 -2.5

12 -3.2 1.3 -2.6 16 -2.7 1.3 -2.1

2 68 -3.0 -2.9

-2.9a 0.5 2.5 -1.2 -1.2

-1.1a 0.1 75 -3.3 -3.4 3.0 -1.1 -1.1 89 -2.7 -2.4 2.4 -1.1 -1.0

3

110 -2.0 -1.7

-2.1b 0.3

1.1 -1.9 -1.6

-2.2a 0.9 117 -2.6 -2.1 1.3 -2.1 -1.6 123 -2.3 -1.6 0.5 -4.5 -3.1 131 -2.3 -1.9 1.4 -1.6 -1.4 138 -2.2 -1.9 1.0 -2.2 -2.0

4 151 -2.2 -1.9

-2.1b 0.3 0.3 -7.7 -6.7

-4.7b 3.0 165 -2.5 -1.7 1.0 -2.5 -1.7

Note: Different superscripts denote significant differences between periods not sharing the same superscript.

Identical superscripts denote no significant difference. Values on period 1 was not included due to different

variation

0

1

2

3

4

5

6

7

8

0 2 4 6 8 10 12

Co

nce

ntr

atio

n (

mg/

L)

Hours

N-NH4

N-NO2

N-NO3

ka kb

Dudy Fredy 35

The specific ammonium conversion rate in period 1 was 2.4±0.2 mg/gVSS.h and then decreased in

period 2 to 1.1±0.1 mg/gVSS.h. The specific ammonium conversion rate of period 3 (2.2±0.9

mg/gVSS.h) seemed higher than period 2 (1.1±0.1 mg/gVSS.h), but the difference was not significant

(P>0.05).

5.2 Nitrogen balance

Table 5.3demonstrates nitrogen balance per cycle in different periods after the ammonium was

completely converted. The nitrified ammonium and the ammonium uptake by nitrifiers were

calculated stoichiometrically by using Equation 4.4. The calculation is based on the amount of nitrate

that was formed in the system during react (aerobic) phase. The ammonium uptake by algae was

calculated from the difference between initial concentrations and the uptake by the nitrifiers.

Table 5.3: Nitrogen balance based on one cycle operation in SBR

Period Day

Initial N-NH4+

concentration (mg/d)

N-NO3-

formed (mg/d)

Daily nitrified N-NH4+ (mg/d)

N-NH4+ uptake by nitrifiers (mg/d)

N-NH4+ uptake by algae (mg/d)

a b c % d avg StdDev % e avg StdDev %

1

4 18.8 14.6 15.0

77

0.4

0.5 0.0 2

3.8

4.8 1.1 23 9 21.2 14.8 15.2 0.4 6.0

12 20.2 15.8 16.3 0.5 3.9 16 21.8 15.8 16.3 0.5 5.5

2 68 27.2 9.5 9.8

48 0.3

0.4 0.1 1 17.4

14.1 5.5 52 75 28.4 11.0 11.3 0.3 17.1 89 24.0 15.7 16.2 0.5 7.8

3

109 26.94 11.3 11.6

41

0.3

0.3 0.0 1

15.4

16.6 1.3 59 117 28.07 11.3 11.6 0.3 16.4 123 27.05 10.7 11.0 0.3 16.0 131 30.40 11.3 11.6 0.3 18.8 138 28.20 11.3 11.7 0.3 16.5

4 151 24.68 11.9 12.3

59 0.4

0.4 0.1 2 12.4

10.2 3.2 41 165 24.72 16.3 16.8 0.5 7.9

Assumption: - no NO3 uptake by algae - no denitrification was occured

Note:

a calculated from measured influent concentration and dilution factor

b taken from the highest concentration in one cycle

c and d are calculated stoichiometrically from Equation 4.4 e = a-c

From Table 5.3, it shows that ammonium uptake by algae is increasing from 4.8±1.1 mg/d in period 1

up to 16.6±1.3 mg/d in period 3, with the assumption that there was no nitrate uptake by algae. On the

36 MSc Thesis

other hand the ammonium that assimilated into nitrifiers is decreasing from 0.5±0.0 mg/d in period 1

to 0.3±0.0 mg/d in period 3.

5.3 Chlorophyll-a concentration

Chlorophyll-a concentration profile of all periods in the SBR is shown in Figure 5.3. A significant

difference between chlorophyll concentration of period 1 and period 2 could be demonstrated

(W=210.5, P<0.01). The chlorophyll-a concentration in period 1 was 19.6±10.4 mg/L then increased

to 28.1±8.1m g/L in period 2 as illustrated in Figure 5.4.

Figure 5.3: Profile of chlorophyll-a concentration in different periods

As shown in Figure 5.4, it is seemed that the chlorophyll of period 3 (7.9±3.0 m/L) was higher than

with period 4 (5.1±2.1 mg/L), but there was not a significant difference (P>0.05). On the other hand if

we look at the chlorophyll-a concentration profile in Figure 5.3, the increasing trend of chlorophyll

concentration in period 4 can be observed.

Figure 5.4: Average chlorophyll-a concentration in different periods

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

40.0

45.0

0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180

Ch

loro

ph

yll-

a co

nce

ntr

atio

n (

mg/

L)

Days

19.6

28.1

7.9 5.1

0.0

5.0

10.0

15.0

20.0

25.0

30.0

35.0

40.0

1 2 3 4

Co

nce

ntr

atio

n (

mg/

L)

Period

Period 1 Period 2 Period 3 Period 4

Dudy Fredy 37

5.4 Suspended solids concentration

TSS and VSS concentration profile of all periods in the SBR are shown in Figure 5.5. The statistical

tests were done to see the differences of VSS average between each period. A significant difference

between VSS of period 1 and period 2 could be demonstrated (W=228.5, P<0.001). The VSS in period

1 was 1.7±0.7 g/L then increased to2.3±0.6 g/L in period 2 as illustrated in Figure 5.5

Figure 5.5: Profile of SS concentration in different periods

The VSS of period 3 (1.0±0.3 m/L) was not significantly different with period 4 (1.0±0.5 m/L)

(P>0.05). However if we look the biomass productivity from Figure 5.6, we can see that in period 4

the productivity is higher than in period 3.

Figure 5.6: Average SS concentration in different periods

0.00

0.50

1.00

1.50

2.00

2.50

3.00

3.50

4.00

4.50

0 10 20 30 40 50 60 70 80 90 100 110 120 130 140 150 160 170 180

SS c

on

cen

trat

ion

(g/

L)

Days TSS VSS

Biomass Productivity

avg Std dev avg Std dev (g/L.d) avg Std dev

1 2.6 0.9 1.7 0.7 0.035 0.65 0.11

2 2.9 0.8 2.3 0.6 0.045 0.82 0.03

3 1.1 0.4 1.0 0.3 0.037 0.90 0.04

4 1.1 0.6 1.0 0.5 0.053 0.91 0.03

PeriodTSS (g/L) VSS (g/L) VSS/TSS

0.0

1.0

2.0

3.0

4.0

1 2 3 4

SS C

on

cen

trat

ipn

(g/

L)

Period

TSS VSS

Period 2 Period 1 Period 3 Period 4

38 MSc Thesis

5.5 Light absorption

The light intensity was measured at the 12 points outer side of the reactor wall, and the average light

intensity was 25.9 µmol/s.m2, as shown in Figure 5.7a. The estimation of light penetration in the

reactor was developed using the Lambert-Beer Law. This is a basis for measuring the amount of light

absorption by the biomass inside the reactor. For simplification, a rectangular shape of reactor can be

assumed as its side view with radius 0.075 m (Figure 5.7b).

The light penetration into the reactor for each period can be seen in Figure 5.8. It can be seen that for

period 1 the light could only penetrate up to 1 cm from the outer wall, while in period 2, 3 and 4 the

light can penetrate up to 2 cm, 4 cm, and centre of the reactor respectively.

Figure 5.7: Light intensity measurement points (a) and a simplified side view of the reactor (b)

Figure 5.8: Estimated light penetration inside reactor

REACTOR

Lamp

D

Lamp

C

Lamp

A

Lamp

B

14.5 cm14.

5 cm

16.0 cm16.0

cm

A

E

C

D

B

F

G

H

I

J

K

L

0.0

4.0

8.0

12.0

16.0

20.0

24.0

28.0

0 0.01 0.02 0.03 0.04 0.05 0.06 0.07 0.08 Ligh

t In

ten

sity

( µ

mo

l/s.

m2

)

Reactor radius (m)

Period 1 Period 2 Period 3 Period 4

LAMP D

LAMP B

LAMP C

LAMP A

r=0.075 m

REACTOR

a b

Dudy Fredy 39

5.6 Biomass composition

The biomass compositions were estimated based on the calculated nitrogen balances in section 5.2 and

a described calculation on section 4.6.3. Figure 5.9 shows that the VSS biomass composition is

consisted mostly by algae. Generally, the percentage of VSS algae in biomass increased over the

period, except in period 4 (50% in period 1, 75% in period 2, 79% in period 3, and 65% in period 4).

Conversely, all the microbes’fraction (nitrifiers, OHOs, and denitrifiers) in biomass had a decreasing

trend, except in period 4. The summary of biomass calculation is presented in Table 5.5.

Figure 5.9: Biomass compositions in each period

5.7 Solids Retention Time (SRT)

The actual SRT was determined based on Equation 4.3, and the results are shown in Table 5.4.

Table 5.4: Actual SRT calculation

Period Vol reactor

(L)

Xr

(g/L) Qw (L/day)

Qs

(L/day) Xs (g/L) SRT (day)

Period 1 1 2.58 0.013 1 0.02 48

Period 2 1 2.86 0.011 1 0.02 52

Period 3 1 1.09 0.029 1 0.01 26

Period 4 1 1.06 0.049 1 0.01 17

Where:

Xr biomass concentration in the reactor (mg/L)

Xs biomass concentration in supernatant (mg/L)

Qw waste discharge flowrate (L/day)

Qs supernatant discharge flowrate (L/day)

0%

20%

40%

60%

80%

100%

1 2 3 4

3.19 1.65 1.18 2.14

39.26 19.75 17.44

27.96

7.28

3.77 2.68

4.89

50.26

74.83 78.7 65.02

Pe

rce

nta

ge V

SS in

bio

mas

s

Period

VSS algae

VSS denitrifiers

VSS OHO

VSS nitrifiers

40

M

Sc T

hesis

Table 5.5: Estimation of biomass composition

Assumption : No NO3 uptake by algae Note: f is calculated stoichiometrically from Denitrification Equation 4.5 g is calculated stoichiometrycally based on NO3 formed in Equation 4.4 h is calculated stochiometrically based on COD removed in Equation 4.6 i is calculated stoichiometrically based on NO3 formed in Equation 4.5 j is calculated stoichiometrically based on NH4 uptake in Equation 4.7 k= g + h + i + j

avg Std Dev avg Std Dev avg Std avg Std Dev

f g h i j k

4 164.93 52.58 3.51 45.28 8.013 40.81 97.61 3.60 46.39 8.21 41.81

9 164.93 52.58 3.56 45.28 8.123 64.59 121.55 2.93 37.25 6.68 53.14

12 164.93 52.58 3.80 45.28 8.672 42.59 100.34 3.79 45.13 8.64 42.44

16 164.93 52.58 3.80 45.28 8.672 59.93 117.69 3.23 38.48 7.37 50.93

54 164.93 52.58 3.77 45.28 8.603 98.19 155.85 2.42 29.05 5.52 63.01

68 118.70 52.58 2.28 26.65 5.208 188.70 222.84 1.02 11.96 2.34 84.68

75 177.91 52.58 2.64 50.51 6.017 185.72 244.89 1.08 20.63 2.46 75.84

89 140.16 52.58 3.78 35.30 8.628 84.66 132.37 2.86 26.67 6.52 63.96

109 131.53 52.58 2.71 31.82 6.175 166.47 207.17 1.31 15.36 2.98 80.35

117 170.94 52.58 2.71 47.70 6.195 178.24 234.85 1.16 20.31 2.64 75.89

123 164.93 52.58 2.57 45.28 5.865 173.93 227.65 1.13 19.89 2.58 76.40

131 164.93 52.58 2.71 45.28 6.175 203.94 258.10 1.05 17.54 2.39 79.02

138 129.24 52.58 2.72 30.90 6.217 179.23 219.07 1.24 14.10 2.84 81.81

4 151 187.05 52.58 2.87 54.20 6.548 134.35 197.96 1.45 27.38 3.31 67.87

165 150.52 52.58 3.92 39.47 8.946 86.00 138.34 2.83 28.53 6.47 62.174.032.14 0.98 27.96 4.89 65.020.82 2.23

2.5378.70

Percentage of algae in

VSS (%)

0.54 39.26 6.96

1.04

7.28

3.77

50.26

74.83

8.71

10.40

2.68

Percentage of

denitrifiers in VSS

(%)

1.24

2.38

0.23

COD

consumed for

denitrification

(mg/d)

VSS

nitrifiers

(mg/d)

Total VSS

(mg/d)

VSS

algae

(mg/d)

1.18

Percentage of

nitrifiers in VSS (%)

Percentage of OHO in

VSS (%)

0.10 17.44 2.73

19.75 7.39

3.19

1.65

3

Period Day

2

1

VSS

denitrifiers

(mg/d)

VSS OHO

(mg/d)

Daily COD

removed

(mg/d)

Dudy Fredy 41

5.8 Oxygen production

The oxygen production by algae in the reactor was estimated according to DO concentration

measurement data. The typical DO profile of one cycle operation in SBR is illustrated in Figure 5.10,

5.11, 5.12, and 5.13.

The calculation procedure was described in section 4.6.4 and the results are summarized in Table 5.6.

Because of the same reactor set-up (stirrer velocity and medium), the oxygen mass transfer coefficient

(KLa) was obtained from previous experiments done by Windaswara (2013).

Figure 5.10: Typical DO profile in one cycle operation in period 1(day 46)

0

2

4

6

8

10

12

14

16

18

20

0 1 2 3 4 5 6 7 8 9 10 11 12

DO

co

nce

ntr

atio

n (

mg/

L)

Time (h)

anoxic aerobic 1 aerobic 2 settling

influent COD source

42 MSc thesis

Figure 5.11: Typical DO profile in one cycle operation in period 2(day 68)

Figure 5.12: Typical DO profile in one cycle operation in period 3 (day 117)

0

2

4

6

8

10

12

14

0 1 2 3 4 5 6 7 8 9 10 11 12

DO

co

nce

ntr

atio

n (

mg/

L)

Time (h)

anoxic 1 aerobic 1 anoxic 2 aerobic 1 settling

influent + COD

influent + COD

0

2

4

6

8

10

12

14

16

18

0 1 2 3 4 5 6 7 8 9 10 11 12

DO

co

nce

ntr

atio

n (

mg/

L)

Time (h)

influent + COD

influent + COD

anoxic 1 aerobic 1 anoxic 2 aerobic 2 settling

Dudy Fredy 43

Figure 5.13: Typical DO profile in one cycle operation in period 4 (day 171)

The statistical tests were done to see the differences of specific oxygen production rate between each

period. A significant difference between the specific oxygen production rate of period 1 and period 2

could be demonstrated (W=128, P<0.001). The specific oxygen production rate in period 1 was

4.1±0.2 mgO2/gVSS.h and then decreased in period 2 to 1.9±0.7 mg O2/gVSS.h.

A Tukey’s HSD Post-hocs test were also be done to determine which period differ from each other.

Similar with the specific ammonium removal rate, it was found that even though the oxygen

production rate of period 3 (6.9±1.8 mgO2/gVSS.h) seemed higher than for period 2 (1.9±0.7 mg

O2/gVSS.h), there was not a significant difference (P>0.05).

Table 5.6: Estimation of oxygen production and consumption rate in the reactor

Period

Oxygen production by algae

Oxygen consumption rate for nitrification

Oxygen consumption rate by OHOs

(mg O2/L.h) (mg

O2/gVSS.h) (mg O2/L.h)

(mg O2/gVSS.h)

(mg O2/L.h) (mg

O2/gVSS.h)

1 9.7±1.0 4.1±0.2 13.4±1.0 5.6±0.3 40.1±5.6 16.9±3.0

2 4.9±1.4a 1.9±0.7a 8.7±1.4 3.4±0.9 79.4±29.6 31.1±12.0

3 7.4±1.7b 6.9±1.8a 11.2±1.7 10.6±3.1 134.6±25.6 129.5±52.2

4 7.8±2.4b 18.2±14.0b 11.6±2.4 26.0±17.9 142.8±41.3 298.9±175.3 Note: Different superscripts denote significant differences between periods not sharing the same superscript.

Identical superscripts denote no significant difference. Values on the period 1 were not included due to different

variation.

0

2

4

6

8

10

12

14

16

18

20

0 1 2 3 4 5 6 7 8 9 10 11 12

DO

co

nce

ntr

atio

n (

mg/

L)

Time (h)

influent + COD

influent + COD

anoxic 1 aerobic 1 anoxic 2 aerobic 2 settling

44 MSc thesis

5.9 Nitrate uptake batch experiment

Based on the measurement results of each one cycle operation, it is known that the nitrate

concentration in settling phase was not decreasing. It means that there was no indication that algae

could assimilate nitrate into its biomass. Therefore to have a better understanding on nitrate uptake by

algae in the reactor, a batch experiment was conducted (see section 4.8).

The batch experiment was conducted within ten days. The refill of NO3- was done on the seventh day.

The nitrate removal rate became higher after the second addition of the NO3-, from 1.3-2.5 mg/L.day

in the first filling to 5.2-5.3 mg/L.day at the second filling as shown in Figure 5.14.

Figure 5.14: Nitrate uptake performances by algae-bacteria consortium

5.10 Microscopic observation

During the experiments, microscopic observation was done weekly. All the microalgae that were

cultured from the start of the experiment still appeared in the reactor at the end of the experiments, as

shown in Figure 5.15-5.16. Figure 5.17 shows an algal-bacterial flocs in the photo-bioreactor.

Figure 5.15: Chlorella sp. and Spirulina sp. (20x magnification)

0.00

2.00

4.00

6.00

8.00

10.00

12.00

14.00

16.00

18.00

20.00

0 1 2 3 4 5 6 7 8 9 10 11

N-N

O3

co

nce

ntr

atio

n (

mg/

L)

Day

0 mg/L of N-NO3

9 mg/L N-NO3

18 mg/L N-NO3

Dudy Fredy 45

Figure 5.16: Scnedesmus sp. and Anabaena sp. (40x magnification)

Figure 5.17: Algal-bacterial flocs

46 MSc thesis

Dudy Fredy 47

6 Discussion

6.1 Ammonium conversion rate

It is hypothesized that the ammonium conversion rate in the algal-bacterial photobioreactor is (i)

proportionally correlated with dissolved oxygen concentration and light penetration into the reactor,

and (ii) inversely correlated with biomass concentration. Coincide with algal-biomass assimilation the

ammonium conversion can be occurred through nitrification (Muñoz et al., 2005; González et al., 2008;

Karya et al., 2013), which its rate is affected by oxygen supply in the reactor. An adequate dissolved

oxygen level is required to maintain a high nitrification rate (Campos et al., 2007). In photosynthetic

oxygenation system, light is a basic energy source for algae to produce oxygen. To enhance

photosynthesis efficiency, an appropriate light intensity should be provided in photobioreactor

(Carvalho et al., 2011). This light intensity depends on its penetration which corresponds with biomass

concentration. The higher biomass concentration exhibit the larger mutual shading of the cell, which

then reduces the penetration depth (Grima et al., 1999).

In this study ammonium conversion rates were calculated based on the slope of ammonium

concentration profile in one cycle operation. It yielded zero order rates as the curves showed linear

line, especially for the first two hours of aerobic phase. Moreover, the nitrification is at maximum rate

(zero-order) when ammonium concentration is higher than its half saturation constant (KS,NH4). It is

reported that half saturation constant for ammonium (KS,NH4) is about 1.2 mg/L (Wolf et al., 2007), so

it is likely that in react phase, the ammonium removal could follow zero order reaction. However as

the ammonium concentration became lower, and then the reaction is changed to a transition between

zero order and first order kinetics. In the normal nitrification process, the ammonium conversion

mostly follows first order kinetics reaction.

From the results of the experiments, the specific ammonium conversion rate in period 1 was higher

than in period 2. This could be due to the limited oxygen concentration in the reactor in period 2. In

period 2, ammonium and COD source are fed simultaneously, induced the competition between OHOs

and nitrifiers, whereas in period 1 such competition did not exist. On the other hand, when the SRT

decreased from 52 days (period 2) to 26 days (period 3), the specific ammonium conversion rate

seemed to increased, from 1.1±0.1 mg/gVSS.h (period 2) and 2.2±0.9 mg/gVSS.h (period 3). However

there was not significant different can be observed between rate in period 2 and period 3 (P>0.05).

But when SRT was decreased again to 17 days (period 4), the specific ammonium conversion rate

increased to 4.7±3.0 mg/gVSS.h. This may related to higher oxygen supply in the reactor during

48 MSc thesis

period 4. And this is consistently supported by the results of oxygen production calculation. In all

periods the oxygen production (volumetric and specific) rates have a similar trend with the specific

ammonium conversion rate. For example, the specific oxygen production rate increased from 6.9±1.8

mgO2/gVSS.h (period 3) to 18.2±14.0 mgO2/gVSS.h (period 4).

However, in this study we could not see the similar trend for volumetric ammonium conversion rate.

This might be due to smaller sample size.

Following the hypothesis, the ammonium conversion rate is also affected by the light penetration into

the reactor and biomass concentration. The calculated estimation of light penetration inside the reactor

shows the similar trend with oxygen production rate (volumetric and specific) and the specific

ammonium conversion rate in all periods. Likewise, the similar trends also is shown from the inverse

measured biomass concentration (TSS/VSS and Chlorophyll-a).

Based on the discussion above, it can be said that the hypothesis may be true on specific ammonium

conversion rate.

6.2 The effects of different operational sequences in SBR

Based on a study by Wu et al (2011), the sequential operation mode in period 1 was applied, and it

supported nitrification and denitrification successfully (Windraswara, 2013). However, it does not

represent the actual wastewater treatment process since the influent of carbon source was fed at a

different time than the ammonium source. Therefore in period 2 to 4, the sequential operation mode

was changed, where the influent of ammonia and carbon source was fed simultaneously.

The results of this study show that ammonium is completely removed, through nitrification and algae

uptake, in all periods. And the sequential operation in period 1 resulted in higher total nitrogen

removal efficiency. The total nitrogen removal efficiency was observed lower in period 2, due to the

fact that the aerobic phase occurred just before the settling phase, which caused the nitrate to appear in

the effluent.

Based on the nitrogen concentration profile in one-cycle operation, the concentration of nitrate during

the settling phase did not change, neither during the react phase. During react phase, even though

ammonium concentration down to zero, but the formed nitrate concentration did not change. This

shows that there was no observable nitrate uptake took place. In contrast, it is reported in literature that

algae can take up the nitrate when all the ammonium has been used up (Borowitzka, n.d.). This may be

because of the nitrate uptake by algae require sufficient time for acclimation and for uptake, to switch

Dudy Fredy 49

from ammonium as nitrogen supply. According to the batch experiment result, the required time to

assimilate about 2 mg N-NO3-/L is 24 hours. This uptake rate could not be observed during the short

duration of the phases.

The operation in period 1 produced an increasing biomass concentration. At the end of period 1, the

competition between heterotrophic microbes and algae for carbon source might have appeared and

resulted in high concentration of nitrate in the effluent. For an optimal denitrification process the

carbon availability should be enough. In this study, it is not only the OHOs and denitrifiers who utilize

the carbon, but maybe also the algae. After double dosage of carbon was applied, it was observed that

the denitrification process performed well; there were no nitrates in the effluent anymore. This

confirmed that denitrification was limited by carbon source, which may have been caused by algae

mixotrophic growth. They can switch their metabolism from autotrophic to heterotrophic in dark

condition. It is in accordance with literature that explained Chlorella sp could grow heterotrophically

and uptake acetate as carbon source (Perez-Garcia et al., 2011).

6.3 The effects of different SRT in SBR

It is reported that the SRT can be effectively control the microorganism’s growth in the reactor

through wastage the mixed liquor (Rittmann and McCarty, 2001). The longer SRT can accommodate a

secure nitrification process. Because nitrifiers are slow growing autotrophic microorganisms, need a

longer time to increase its capacity for removing ammonium.

Generally, the ammonium removal efficiency and nitrogen removal were not different for each SRT

(P>0.05). At all periods, the ammonium was removed completely through nitrification and biomass

accumulation. The percentage of ammonium removal pathway in the reactor can be seen through

result of nitrogen balance calculation. Nitrification was responsible for 48% (period 2), 41% (period 3),

and 59% (period 4) of ammonium removal, whereas algae biomass accumulation was responsible for

52%, 59%, and 41%, respectively. No statistical differences were observed for the nitrogen balances

calculation for different periods.

Based on the results, with decreasing SRT the biomass concentration (both chlorophyll-a and

suspended solids) is significantly decreased from period 2 to period 3. However, the biomass

concentration in period 4 (SRT 17d) was not significantly different with that in period 3 (SRT 26 d)

(P>0.05). It might be that in period 4 the decreased SRT did not affect in decreased biomass but it

reaches some point where the biomass started to increase again. It is probably due to better light

intensity and oxygen production.

50 MSc thesis

The biomass productivity in lower SRT (period 4) (0.053 gVSS/L.d) was higher than in period 3(0.037

gVSS/L.d). These results are in accordance with a study by Valigore et al,. (2012), who identified that

a lower SRT could enhanced the biomass productivity in algal-bacterial biomass grown on primary

treated wastewater.

Moreover according to Janssen and Lamers (2013), the algal photobioreactor volumetric productivity

(rux) is a function of biomass concentration (Cx) and the specific growth rate of microalgae (µ), which

is light-limited growth. And the relation can be plotted as shown in Figure 6.1.The photobioreactor

will operate at its highest productivity at a certain biomass concentration (Cx,opt). In this research

lowering biomass concentration via lowering the SRT could decrease the biomass concentration, and

therefore might increase the photobioreactor productivity.

Figure 6.1: Volumetric productivity of a photobioreactor rUx as a function of biomass concentration Cx

Source: (Janssen and Lamers, 2013)

On the other hand, from nitrifiers kinetic growth equation below:

(6.1)

the ammonium conversion rate (dNa/dt) is proportional with biomass concentration XBA. Nitrifiers as

slow growing autotrophic microorganisms need a higher biomass concentration to increase its capacity

for removing ammonium. But the higher the concentration of nitrifiers, the more shading they are

causing, which reduces the photosynthesis by algae and the production of oxygen in the reactor.

Dudy Fredy 51

From above discussion, it can be identified that reducing the SRT may increase oxygen production

rate, and may increase ammonium conversion rate, even only a small effect. It can also be identified

that reducing SRT may increase biomass productivity, but that will lead to larger mutual shading. On

the other hand a long SRT can guarantee the nitrification occurred completely and produce more

stable sludge. But also could require larger volume of reactor due to accumulation of fast growing

heterotrophs and inert suspended solids.

A decision on what SRT should be applied on photo-activated sludge should consider on an optimum

biomass concentration that can provide enough oxygen for nitrification but do not lead to larger

mutual shading.

6.4 Comparison with other algal-bacterial photo-bioreactor

The comparison with other researches can be summarized in Table 6.1.

Table 6.1: Comparison the result with other research

Research /Refference

Max ammonium

conversion rates

(mg N/L.h)

Oxygen

production

rate (kg

O2/m3.d)

Biomass

productivity

(gVSS/L.d)

Ammonium

removal via

biomass

accumulation(%)

This research 3.4 0.3 0.03-0.05 23-59

(Karya et al., 2013) 7.7 0.46 - 15-19

HRAP (Muñoz and

Guieysse, 2006) - 0.3-0.38 - -

Enclosed Photobioreactor

(Muñoz and Guieysse,

2006)

- 1.8-8.3 - -

(Su et al., 2011) - - - 44

MaB-flocs (Van Den

Hende et al., 2011) - - 0.15-0.18 -

The observed maximum ammonium conversion rate was 3.4 mgN-NH4/L.h. However, this values are

lower than reported in previous study (7.7 mgN-NH4/L.h) (Karya et al., 2013). The low value is

probably due to lower light intensity applied. And competition with hetereotrophs for oxygen due to

COD source was fed in this study.

The maximum observed oxygen production rate occurred in period 4 (0.30 kgO2/m3.day). This value

was lower if we compared with A previous study by Karya et al. (2013) (0.46 kgO2/m3.day), and far

below the rates that had been produced by enclosed photo-bioreactors (1.8-8.3 kgO2/m3.day) (Muñoz

and Guieysse, 2006). However, the oxygen production rate in this study was similar with the oxygen

52 MSc thesis

production by HRAP (0.30-0.38 kgO2/m3.day) (Muñoz and Guieysse, 2006). This lower oxygen

production in this study was probably mainly caused by a lower applied light intensity.

The results of nitrogen balance calculation showed that on average the ammonium removal was done

via nitrification (41-77%) and algae uptake (23-59%). The effect of different SRT could not be

observed due to insignificant different between periods (P>0.05). These results are in accordance with

the study that used municipal wastewater, where the biomass accumulation was 44% (Su et al., 2011).

Moreover, the biomass accumulation in this study is higher compared with previous study by Karya et

al., (2013) about 15-19%.

The biomass productivity in this study ranged between 0.03-0.05 gVSS/L.day. It is lower compared

with microalgal-bacterial flocs (MaB-flocs) from other lab scale photo-bioreactor which has

productivity around 0.15-0.18 gVSS/L.day (Van Den Hende et al., 2011).

Based on nitrogen mass balances, the fraction of biomass was calculated. It was found that the

biomass of is consisted of nitrifiers (1.2-3.4%), algae (48-79%) and the heteretrophs (19-45%). The

low percentage of nitrifiers in the biomass is probably due to nitrification inhibition by algae, where

the both compete for the same carbon and nitrogen source (Choi et al., 2010).

6.5 Development of biofilm in the reactor

The development of biofilm attached on the reactor’s wall became more intense after about 100 days

of operation. In this study, the biofilm occurrence was considered as a negative factor. Since it reduced

the light penetration into reactor and created unequal biomass concentration within the reactor. The

formation of biofilm was influenced by many factors, such as extra polymeric substances (EPS)

produced by bacteria and microalgae species selection (Irving and Allen, 2011). The microscopic

observation was done, the species which was dominant sticking on the wall is the thread forming

microalgae. This species was the filamentous algae, and it may be the Spirulina sp. (Figure 6.2).

The reason why the algae have a tendency to adhere on the wall of the reactor is still unclear. It may

be due to its motility for a better light catchment, or to have a better nutrient source. The coexistence

with microbes which also produce EPS, might be another factor that can increase the tendency of

microalgae to attach on any surfaces.

Dudy Fredy 53

Figure 6.2: Attached thread-former species of microalgae in reactor’s wall

(a=10x magnification, b= 20x magnification, and c=40x magnification)

Some possible measures to reduce the tendency of biofilm formation are as follows:

a. Applying higher stirring velocity without destructing the biomass.

b. Microalgae species selection.

c. Consideration on other reactor design for easier maintenance work, such as flat panel type of

reactor.

6.6 Light regime in photobioractor

The light regime inside a reactor can be characterized with a photic zone of intense light at the reactor

surface and a dark zone in the interior of the reactor. A photic zone is defined as the depth at which 90%

of the incoming photon flux is absorbed, and called the light penetration depth (dp) (Janssen et al.,

2003). The relative photic volume (ε) is defined as the ratio of the volume of the photic zone over the

reactor volume. The estimated light regime for light absorption by chlorophyll at wavelength 550 nm

inside the reactor is presented in Figure 5.9.

Figure 6.3: Light fraction as function of the chlorophyll-a concentration in the reactor

As can be seen from Figure 5.2, the higher concentration of biomass it can reduce the light fraction

absorbed in the reactor. The light can only penetrate into the interior reactor when the cholorophyll

0.0

0.2

0.4

0.6

0.8

1.0

0.00 5.00 10.00 15.00 20.00 25.00 30.00 The

re

lati

ve p

ho

tic

volu

me

)

Chlorophyll-a (mg/L)

a c b

54 MSc thesis

less than 5 mg/L. This is a rough estimation, but it may indicate that on period 1 or 2, the light hardly

penetrates into the reactor. This may create dark zone in the reactor, which may lead to anoxic

condition.

6.7 Denitrification

Figure 6.4 shows the typical nitrogenous concentration profile in one cycle of SBR operation. It can be

seen that denitrification was occurred at both anoxic phases. Because of dark period and process of

organic degradation by OHOs, the DO concentration dropped to near zero. This condition is favorable

for denitrification to be occurred. It also can be seen that the nitrate formed by nitrification was not

denitrified. This is becaused of no more COD appear in the reactor, which all used by OHOs.

Figure 6.4: Typical nitrogen concentration profile within a cycle (Day 117)

Nitrogen mass balance calculation was done and based on nitrate formation by nitrification process. In

calculation, the difference between nitrogen from influent ammonium with nitrate formed was

assumed to be assimilated by algae. This nitrogen uptake by algae may be over estimated, the error

could arise due to its purpose to close the nitrogen balance. Beside algae uptake another possibility for

balancing the nitrogen is denitrification. As discussed on previous section that the dark zone could be

appear inside the reactor, due to shorter light penetration. If it is the case, the algae-bacteria flocs or

aggregates would have an anoxic zone and perform denitrification.

0.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

9.0

0 2 4 6 8 10 12

N-NH4 (mg/L)

N-NO2 (mg/L)

N-NO3 (mg/L)

Time (h)

Co

nce

ntr

atio

n (

mg/

L)

denitrification denitrification

Nitrogen assimilated by algae and possibly denitrification

Dudy Fredy 55

7 Conclusion and recommendations

7.1 Conclusion

The present study has shown that SRT has only limited effect on the nitrification process and algae

uptake by algal-bacterial biomass in a photo-bioreactor. The overall ammonium removal rates only

varies from 2.1 to 2.9 mgN-NH4/L.h , while SRT varies from 17 to 52 days. The maximum oxygen

production was occurred in SRT 17 days at a rate 0.3 mgO2/m3.day.

7.2 Recommendations

The following experimental conditions are advised in order to achieve a better understanding of algal-

bacterial system in Sequential Batch Photo-bioreactor:

1. Investigating the denitrification performance, if applicable determination of nitrogen gas

formation

2. Applying higher light intensity

3. Investigating higher ammonium loading rate

4. Application of actual wastewater

5. Investigating the effect of pH

56 MSc thesis

Dudy Fredy 57

References

Abeliovich, A., Vonshak, A.W.A., 1993. FACTORS INHIBITING NITRIFICATION OF AMMONIA

IN DEEP WASTEWATER RESERVOIRS. Water Research 27, 1585–1590.

Babu, M. a., Hes, E.M. a., van der Steen, N.P., Hooijmans, C.M., Gijzen, H.J., 2010. Nitrification rates

of algal–bacterial biofilms in wastewater stabilization ponds under light and dark conditions.

Ecological Engineering 36, 1741–1746.

Becker, E.W., 1994. Microalgae : Biotechnology and Microbiology. Cambridge University Press,,

Cambridge ;

Boelee, N.C., Temmink, H., Janssen, M., Buisman, C.J.N., Wijffels, R.H., 2011. Nitrogen and

phosphorus removal from municipal wastewater effluent using microalgal biofilms. Water

research 45, 5925–33.

Boelee, N.C., Temmink, H., Janssen, M., Buisman, C.J.N., Wijffels, R.H., 2012. Scenario Analysis of

Nutrient Removal from Municipal Wastewater by Microalgal Biofilms. Water 4, 460–473.

Borowitzka, M., n.d. Topic 8 : Algal Culture and Biotechnology.

Campos, J.L., Garrido, J.M., Mosquera-Corral, a., Méndez, R., 2007. Stability of a nitrifying activated

sludge reactor. Biochemical Engineering Journal 35, 87–92.

Carvalho, A.P., Silva, S.O., Baptista, J.M., Malcata, F.X., 2011. Light requirements in microalgal

photobioreactors: an overview of biophotonic aspects. Applied microbiology and biotechnology

89, 1275–88.

Chernicharo, C. a. L., 2006. Post-Treatment Options for the Anaerobic Treatment of Domestic

Wastewater. Reviews in Environmental Science and Bio/Technology 5, 73–92.

Chiu, S.-Y., Kao, C.-Y., Chen, C.-H., Kuan, T.-C., Ong, S.-C., Lin, C.-S., 2008. Reduction of CO2 by

a high-density culture of Chlorella sp. in a semicontinuous photobioreactor. Bioresource

technology 99, 3389–96.

Chiu, Y., Chung, M., 2003. Determination of optimal COD = nitrate ratio for biological denitriÿcation

51, 43–49.

Choi, O., Das, A., Yu, C.-P., Hu, Z., 2010. Nitrifying bacterial growth inhibition in the presence of

algae and cyanobacteria. Biotechnology and bioengineering 107, 1004–11.

Chong, S., Sen, T.K., Kayaalp, A., Ang, H.M., 2012. The performance enhancements of upflow

anaerobic sludge blanket (UASB) reactors for domestic sludge treatment--a state-of-the-art

review. Water research 46, 3434–70.

Christenson, L., Sims, R., 2011. Production and harvesting of microalgae for wastewater treatment,

biofuels, and bioproducts. Biotechnology advances 29, 686–702.

Craggs, R., Sutherland, D., Campbell, H., 2012. Hectare-scale demonstration of high rate algal ponds

for enhanced wastewater treatment and biofuel production. Journal of Applied Phycology 24,

329–337.

58 MSc thesis

de-Bashan, L.E., Moreno, M., Hernandez, J.-P., Bashan, Y., 2002. Removal of ammonium and

phosphorus ions from synthetic wastewater by the microalgae Chlorella vulgaris coimmobilized

in alginate beads with the microalgae growth-promoting bacterium Azospirillum brasilense.

Water research 36, 2941–8.

Ekama, G. a, 2010. The role and control of sludge age in biological nutrient removal activated sludge

systems. Water science and technology : a journal of the International Association on Water

Pollution Research 61, 1645–52.

Ekama, G.A., Wentzel, M.C., 2008. Biological Wastewater Treatment: Principles, Modelling and

Design. IWA Publishing.

Fuentes, M.M.R., Garcı, J.L., Acie, F.G., 1999. Outdoor continuous culture of Porphyridium cruentum

in a tubular photobioreactor : quantitative analysis of the daily cyclic variation of culture

parameters 70, 271–288.

Fukami, K., Nishijima, T., Ishida, Y., 1997. Stimulative and inhibitory effects of bacteria on the

growth of microalgae 185–191.

Gerardi, M.H., 2002. Nitrification and Denitrification in Activated Sludge Process. John Wiley &

Sons, Inc.

Gomec, C.Y., 2010. High-rate anaerobic treatment of domestic wastewater at ambient operating

temperatures: A review on benefits and drawbacks. Journal of environmental science and health.

Part A, Toxic/hazardous substances & environmental engineering 45, 1169–84.

González, C., Marciniak, J., Villaverde, S., García-Encina, P. a, Muñoz, R., 2008. Microalgae-based

processes for the biodegradation of pretreated piggery wastewaters. Applied microbiology and

biotechnology 80, 891–8.

Grima, E.M., Acie, F.G., Chisti, Y., 1999. Photobioreactors : light regime , mass transfer , and scaleup

70, 231–247.

Grunditz, C., Dalhammar, G., 2001. Development of nitrification inhibition assays using pure cultures

of Nitrosomonas and Nitrobacter. Water research 35, 433–40.

Guerrero, M., Jones, R., 1996. Photoinhibition of marine nitrifying bacteria. I. Wavelength-dependent

response. Marine Ecology Progress Series 141, 183–192.

Henze, M., 2008. Biological wastewater treatment: principles, modelling and design.

Hooper, a B., Terry, K.R., 1973. Specific inhibitors of ammonia oxidation in Nitrosomonas. Journal of

bacteriology 115, 480–5.

Hsieh, C.-H., Wu, W.-T., 2009. A novel photobioreactor with transparent rectangular chambers for

cultivation of microalgae. Biochemical Engineering Journal 46, 300–305.

Huang, S., Chen, C., Yang, X., Wu, Q., Zhang, R., 2011. Distribution of typical denitrifying functional

genes and diversity of the nirS-encoding bacterial community related to environmental

characteristics of river sediments. Biogeosciences 8, 3041–3051.

Irving, T.E., Allen, D.G., 2011. Species and material considerations in the formation and development

of microalgal biofilms. Applied microbiology and biotechnology 92, 283–94.

Dudy Fredy 59

Janssen, M., Lamers, P., 2013. Microalgae Biotechnology. Wageningen.

Janssen, M., Tramper, J., Mur, L.R., Wijffels, R.H., 2003. Enclosed outdoor photobioreactors: light

regime, photosynthetic efficiency, scale-up, and future prospects. Biotechnology and

bioengineering 81, 193–210.

Jiménez, E., Giménez, J.B., Ruano, M. V, Ferrer, J., Serralta, J., 2011. Effect of pH and nitrite

concentration on nitrite oxidation rate. Bioresource technology 102, 8741–7.

Jiménez, E., Giménez, J.B., Seco, a, Ferrer, J., Serralta, J., 2012. Effect of pH, substrate and free

nitrous acid concentrations on ammonium oxidation rate. Bioresource technology 124, 478–84.

Kaplan, D., Wilhelm, R., Abeliovich, A., 1998. Interdependent environmental factors controlling

nitrification in waters. Water Science & Technology 42, 167–172.

Karya, N.G. a I., van der Steen, N.P., Lens, P.N.L., 2013. Photo-oxygenation to support nitrification in

an algal-bacterial consortium treating artificial wastewater. Bioresource technology 134, 244–50.

Khan, A.A., Gaur, R.Z., Tyagi, V.K., Khursheed, A., Lew, B., Mehrotra, I., Kazmi, a. a., 2011.

Sustainable options of post treatment of UASB effluent treating sewage: A review. Resources,

Conservation and Recycling 55, 1232–1251.

Komorowska-Kaufman, M., Majcherek, H., Klaczyński, E., 2006. Factors affecting the biological

nitrogen removal from wastewater. Process Biochemistry 41, 1015–1021.

Kujawa, K., Klapwijk, B., 1999. A method to estimate denitrification potential for predenitrification

systems using NUR batch test. Water Research 33, 2291–2300.

Larsdotter, K., 2006. Waste Water treatment With microalgae – a literature review. Vatten 62, 31–38.

Lee, J.-K., Choi, C.-K., Lee, K.-H., Yim, S.-B., 2008. Mass balance of nitrogen, and estimates of

COD, nitrogen and phosphorus used in microbial synthesis as a function of sludge retention time

in a sequencing batch reactor system. Bioresource technology 99, 7788–96.

Lee, S., Cho, K., Lim, J., Kim, W., Hwang, S., 2011. Acclimation and activity of ammonia-oxidizing

bacteria with respect to variations in zinc concentration, temperature, and microbial population.

Bioresource technology 102, 4196–203.

Liao, B.Q., Droppo, I.G., Leppard, G.G., Liss, S.N., 2006. Effect of solids retention time on structure

and characteristics of sludge flocs in sequencing batch reactors. Water research 40, 2583–91.

Liu, G., Wang, J., 2012. Probing the stoichiometry of the nitrification process using the respirometric

approach. Water research 46, 5954–62.

Mahvi, A.H., 2008. SEQUENCING BATCH REACTOR: A PROMISING TECHNOLOGY IN

WASTEWATER TREATMENT. Iran J. Environ. Health. Sci. Eng 5, 79–90.

Mara, D.D., 2003. Wastewater Treatment in Developing Countries.

Markou, G., Georgakakis, D., 2011. Cultivation of filamentous cyanobacteria (blue-green algae) in

agro-industrial wastes and wastewaters: A review. Applied Energy 88, 3389–3401.

60 MSc thesis

Mata, T.M., Martins, A. a., Caetano, N.S., 2010. Microalgae for biodiesel production and other

applications: A review. Renewable and Sustainable Energy Reviews 14, 217–232.

Mateju, V., cizinska, SKrejci, J., Janoch, T., 1992. Biological water denitrification-- A review.

Enzyme and Microbial Technology 14.

Moheimani, N.R., Borowitzka, M.A., 2007. Limits to Productivity of the Alga Pleurochrysis carterae (

Haptophyta ) Grown in Outdoor Raceway Ponds 96, 27–36.

Muñoz, R., Guieysse, B., 2006. Algal-bacterial processes for the treatment of hazardous contaminants:

a review. Water research 40, 2799–815.

Muñoz, R., Jacinto, M., Guieysse, B., Mattiasson, B., 2005. Combined carbon and nitrogen removal

from acetonitrile using algal-bacterial bioreactors. Applied microbiology and biotechnology 67,

699–707.

Muñoz, R., Köllner, C., Guieysse, B., Mattiasson, B., 2004. Photosynthetically oxygenated salicylate

biodegradation in a continuous stirred tank photobioreactor. Biotechnology and bioengineering

87, 797–803.

Ogbonna, J.C., Tanaka, H., 2000. Light requirement and photosynthetic cell cultivation –

Development of processes for efficient light utilization in photobioreactors. Journal of Applied

Phycology 12, 207–218.

Oswald, W.J., 2003. My sixty years in applied algology. Journal of Applied Phycology 15, 99–106.

Oswald, W.J., Gotaas, H.B., 1957. Photosynthesis in sewage treatment. American Society of Civil

Engineers Transactions.

Park, J.B.K., Craggs, R.J., Shilton, a N., 2011. Wastewater treatment high rate algal ponds for biofuel

production. Bioresource technology 102, 35–42.

Perez-Garcia, O., De-Bashan, L.E., Hernandez, J.-P., Bashan, Y., 2010. Efficiency of Growth and

Nutrient Uptake From Wastewater By Heterotrophic, Autotrophic, and Mixotrophic Cultivation

of Chlorella Vulgaris Immobilized With Azospirillum Brasilense1. Journal of Phycology 46,

800–812.

Perez-Garcia, O., Escalante, F.M.E., de-Bashan, L.E., Bashan, Y., 2011. Heterotrophic cultures of

microalgae: metabolism and potential products. Water research 45, 11–36.

Posten, C., 2009. Design principles of photo-bioreactors for cultivation of microalgae. Engineering in

Life Sciences 9, 165–177.

Puig, S., Corominas, L., Colomer, J., Balaguer, M.D., Colprim, J., 2005. On-line Oxygen Uptake Rate

as a new tool for monitoring and controlling the SBR process, in: European Symposium on

Computer Aided Process Engineering-15.

Rittmann, B., McCarty, P., 2001. Environmental biotechnology: principles and applications. McGraw-

Hill, Boston.

Rodríguez, D.C., Pino, N., Peñuela, G., 2011. Monitoring the removal of nitrogen by applying a

nitrification-denitrification process in a Sequencing Batch Reactor (SBR). Bioresource

technology 102, 2316–21.

Dudy Fredy 61

Rodríguez, D.C., Ramírez, O., Mesa, G.P., 2011. Behavior of nitrifying and denitrifying bacteria in a

sequencing batch reactor for the removal of ammoniacal nitrogen and organic matter.

Desalination 273, 447–452.

Ryu, H.-D., Lee, S.-I., 2009. Comparison of 4-Stage Biological Aerated Filter (BAF) with MLE

Process in Nitrogen Removal from Low Carbon-to-Nitrogen Wastewater. Environmental

Engineering Science 26, 163–170.

Safonova, E., Kvitko, K.V., Iankevitch, M.I., Surgko, L.F., Afti, I. a., Reisser, W., 2004. Biotreatment

of Industrial Wastewater by Selected Algal-Bacterial Consortia. Engineering in Life Sciences 4,

347–353.

Salem, S., Moussa, M.S., Loosdrecht, M.C.M. Van, 2006. Determination of the Decay Rate of

Nitrifying Bacteria. Biotechnology and Bioengineering 94, 252–262.

Schramm, A., Santegoeds, C.M., Nielsen, H.K., Ploug, H., Wagner, M., Pribyl, M., Wanner, J.,

Amann, R., Beer, D. De, 1999. On the Occurrence of Anoxic Microniches , Denitrification , and

Sulfate Reduction in Aerated Activated Sludge On the Occurrence of Anoxic Microniches ,

Denitrification , and Sulfate Reduction in Aerated Activated Sludge.

Schumacher, G., Blume, T., Sekoulov, I., 2003. Bacteria reduction and nutrient removal in small

wastewater treatment plants by an algal biofilm. Water science and technology : a journal of the

International Association on Water Pollution Research 47, 195–202.

Seghezzo, L., Zeeman, G., Liel, J.B. Van, Hamelers, H.V.M., Lettinga, G., 1998. A REVIEW : THE

ANAEROBIC TREATMENT OF SEWAGE IN UASB AND EGSB REACTORS. Bioresource

technology 65, 175–190.

Singh, M., Srivastava, R.K., 2011. Sequencing batch reactor technology for biological wastewater

treatment : a review. Asia-Pacific Journal of Chemical Engineering 6, 3–13.

Sinha, B., Annachhatre, A.P., 2006. Partial nitrification—operational parameters and microorganisms

involved. Reviews in Environmental Science and Bio/Technology 6, 285–313.

Stanier, R.Y., Kunisawa, R., Mandel, M., Cohen-Bazire, G., 1971. Purification and properties of

unicellular blue-green algae (order Chroococcales). Bacteriological reviews 35, 171–205.

Su, Y., Mennerich, A., Urban, B., 2011. Municipal wastewater treatment and biomass accumulation

with a wastewater-born and settleable algal-bacterial culture. Water research 45, 3351–8.

Subashchandrabose, S.R., Ramakrishnan, B., Megharaj, M., Venkateswarlu, K., Naidu, R., 2011.

Consortia of cyanobacteria/microalgae and bacteria: biotechnological potential. Biotechnology

advances 29, 896–907.

Tam, N., Wong, Y., Leung, G., 1992. Effect of exogenous carbon sources on removal of inorganic

nutrient by the nitrification-denitrification process. Water Research 26, 1229–1236.

Valigore, J.M., Gostomski, P. a, Wareham, D.G., O’Sullivan, A.D., 2012. Effects of hydraulic and

solids retention times on productivity and settleability of microbial (microalgal-bacterial)

biomass grown on primary treated wastewater as a biofuel feedstock. Water research 46, 2957–

64.

62 MSc thesis

Van Den Hende, S., Vervaeren, H., Desmet, S., Boon, N., 2011. Bioflocculation of microalgae and

bacteria combined with flue gas to improve sewage treatment. New biotechnology 29, 23–31.

Vanzella, A., Guerrero, M.A., Jones, R.D., 1989. Effect of CO and light on ammonium and nitrite

oxidation by chemolithotrophic bacteria. Marine Ecology Progress Series 57, 69–76.

Windraswara, R., 2013. Nitrification and denitrification by an algal-bacterial consortium in a photo-

SBR.

Wolf, G., Picioreanu, C., Loosdrecht, M.C.M. Van, 2007. Kinetic Modeling of Phototrophic Biofilms :

The PHOBIA Model 97, 1064–1079.

Wolfaardt, G.M., Lawrence, J.R., Robarts, R.D., Caldwell, D.E., 1994. The role of interactions, sessile

growth and nutrient amendments on the degradative efficiency of a microbial consortium.

Canadian journal of microbiology 40, 331–40.

Wu, C., Peng, Y., Wang, S., Li, X., Wang, R., 2011. Effect of Sludge Retention Time on Nitrite

Accumulation in Real-time Control Biological Nitrogen Removal Sequencing Batch Reactor.

Chinese Journal of Chemical Engineering 19, 512–517.

Appendix A

A.1 Daily Nitrogenous concentration

Period Date Day

Influent concentration

(mg/l)

Effluent concentration (mg/l)

N-NH4+ N-NH4

+ N-NO2-

N-NO3

-

1

18/02/13 1 22.71 0.00 1.20 0.00

20/02/13 3 23.45 0.00 0.00 0.00

21/02/13 4 23.22 0.00 0.00 0.00

22/02/13 5 24.77 0.00 0.00 0.00

25/02/13 8 24.05 0.00 0.00 0.01

26/02/13 9 24.48 0.00 0.00 0.26

27/02/13 10 23.70 0.00 0.00 0.14

28/02/13 11 24.27 0.00 0.00 0.20

1/03/13 12 24.98 0.00 0.00 0.00

4/03/13 15 25.30 0.08 0.00 0.00

5/03/13 16 25.37 0.10 0.00 0.00

6/03/13 17 25.10 0.10 0.00 0.00

7/03/13 18 25.37 0.08 0.00 0.00

11/03/13 22 22.20 0.00 0.00 0.00

12/03/13 23 21.48 0.00 0.00 0.00

13/03/13 24 22.17 0.00 0.00 0.00

18/03/13 29 23.00 0.00 0.00 0.00

19/03/13 30 25.39 0.00 0.00 0.00

20/03/13 31 23.32 0.00 0.00 0.00

21/03/13 32 25.07 0.00 0.00 0.00

28/03/13 39 21.88 0.00 0.00 0.00

2/04/13 44 21.45 0.00 0.00 6.22

3/04/13 45 23.40 0.00 0.00 7.44

4/04/13 46 23.85 0.00 0.00 5.73

5/04/13 47 22.91 0.00 0.00 7.82

8/04/13 50 21.32 0.00 0.00 6.27

9/04/13 51 21.70 0.00 0.00 11.54

10/04/13 52 22.95 0.00 0.06 8.34

11/04/13 53 26.18 0.00 0.01 0.62

12/04/13 54 25.17 0.00 0.00 0.35

15/04/13 57 25.82 0.00 0.00 0.00

16/04/13 58 25.53 0.00 0.00 0.00

17/04/13 59 24.05 0.00 0.00 0.00

18/04/13 60 24.12 0.00 0.00 0.00

19/04/13 61 24.41 0.00 0.00 1.33

22/04/13 64 22.93 0.00 0.00 1.34

23/04/13 65 25.57 0.00 0.00 0.40 24/04/13 66 25.21 0.00 0.00 1.79 25/04/13 67 24.70 0.00 0.00 5.88

2

26/04/13 68 23.30 0.00 0.00 2.19 29/04/13 71 22.96 0.00 0.00 1.59 1/05/13 73 23.82 0.00 0.00 2.66 2/05/13 74 24.26 0.00 0.00 2.57 3/05/13 75 24.36 0.00 0.00 2.57 6/05/13 78 22.50 0.00 0.00 2.68 7/05/13 79 22.93 0.00 0.00 16.36 8/05/13 80 23.18 0.00 0.00 18.22

10/05/13 82 23.43 0.00 0.00 19.75

13/05/13 85 21.71 0.00 0.00 2.07 14/05/13 86 21.32 0.00 0.00 7.23 15/05/13 87 22.36 0.00 0.00 13.09 16/05/13 88 19.89 0.00 0.00 4.75

17/05/13 89 20.57 0.00 0.00 4.59 21/05/13 93 20.55 0.00 0.00 8.36 22/05/13 94 21.12 0.00 0.00 3.32 23/05/13 95 20.88 0.00 0.00 10.05 24/05/13 96 20.58 0.00 0.00 8.81 28/05/13 100 22.95 0.00 0.00 0.60 4/06/13 107 21.28 0.00 0.02 2.06 5/06/13 108 21.68 0.00 0.02 2.23

3

7/06/13 110 21.22 0.00 0.01 2.24

10/06/13 113 22.88 0.00 0.03 2.25

11/06/13 114 23.38 0.00 0.03 2.70

12/06/13 115 24.97 0.00 0.03 2.20

13/06/13 116 23.78 0.00 0.03 2.42

14/06/13 117 23.81 0.00 0.01 2.02

17/06/13 120 23.42 0.00 0.03 2.56

18/06/13 121 22.96 0.00 0.01 3.31

19/06/13 122 23.38 0.00 0.01 1.92

20/06/13 123 22.89 0.00 0.03 2.53

21/06/13 124 20.58 0.00 0.00 2.61

24/06/13 127 20.82 0.00 0.00 1.89

25/06/13 128 24.73 0.00 0.00 2.49

26/06/13 129 24.52 0.00 0.00 2.17

27/06/13 130 24.39 0.00 0.01 2.28

28/06/13 131 24.92 0.00 0.01 2.56

1/07/13 134 22.59 0.00 0.01 2.17

2/07/13 135 23.59 0.00 0.01 2.97

3/07/13 136 23.59 0.00 0.01 2.36

4/07/13 137 22.76 0.36 0.00 2.09

5/07/13 138 23.65 0.14 0.01 2.60

4

8/07/13 141 22.26 0.81 0.01 2.00

9/07/13 142 22.47 0.22 0.01 2.60

10/07/13 143 21.55 0.14 0.01 1.97

11/07/13 144 21.14 0.14 0.01 1.90

15/07/13 148 19.52 0.00 0.26 0.00

16/07/13 149 22.84 0.00 0.28 0.00

17/07/13 150 20.65 0.00 0.22 0.00

18/07/13 151 18.90 0.00 0.22 1.51

19/07/13 152 18.09 0.00 0.30 1.33

22/07/13 155 19.89 0.00 0.22 1.01

23/07/13 156 21.37 0.00 0.26 1.17

24/07/13 157 22.52 0.00 0.18 1.26

1/08/13 165 21.19 0.00 3.87 0.78

5/08/13 169 21.10 0.00 2.08 0.74

6/08/13 170 20.34 0.00 3.06 0.85

A.2 Calculated ammonium conversion rate

Figure A.1: Ammonium conversion rate calculation in period 1 day 4

Figure A.2: Ammonium conversion rate calculation in period 1 day 9

21 February 2013 within a cycle (Day 4)Time (h) N-NH4 N-NO2 N-NO3

(mg/L) (mg/L) (mg/L)

0.0 12.9 0.0 0.0

1.0 7.5 0.1 0.7

1.5 6.2 0.2 1.7

2.5 5.4 0.4 2.1

3.0 4.2 0.5 2.9

4.0 1.7 0.9 4.6

5.0 0.3 0.0 7.2

6.0 0.2 0.0 7.3

7.0 0.3 0.0 7.2

8.0 0.3 1.7 1.4

9.0 0.3 0.0 0.0

10.0 0.2 0.0 0.0

11.0 0.2 0.0 0.0

12.0 0.2 0.0 0.0

y = -2.4982x + 11.318R² = 0.9084

0.0

2.0

4.0

6.0

8.0

10.0

12.0

14.0

0 2 4 6 8 10 12

N-NH4

N-NO2

N-NO3

Linear (Series1)

Time (h)

Co

nce

ntr

atio

n (m

g/L)

26 February 2013 within a cycle (Day 9)Time (h) N-NH4 N-NO2 N-NO3

(mg/L) (mg/L) (mg/L)

0 11.8 0.0 0.0

0.2 10.6 0.0 0.1

1.3 7.1 0.2 1.0

2.3 4.4 0.4 2.1

3.3 1.9 0.7 4.5

4.3 0.2 0.1 6.8

5.3 0.2 0.0 7.4

6.3 0.2 0.0 7.4

7.6 0.2 0.0 7.3

7.8 0.2 2.4 1.8

8.3 0.2 0.0 0.0

9.3 0.2 0.0 0.0

10.3 0.2 0.0 0.0

12.0 0.2 0.0 0.0

y = -2.7084x + 11.121R² = 0.9853

0.0

2.0

4.0

6.0

8.0

10.0

12.0

14.0

0 2 4 6 8 10 12

N-NH4 (mg/L)

N-NO2 (mg/L)

N-NO3 (mg/L)

Linear (N-NH4)

Time (h)

Co

nce

ntr

atio

n(m

g/L)

Figure A.3: Ammonium conversion rate calculation in period 1 day 12

Figure A.4: Ammonium conversion rate calculation in period 1 day 16

1 March within a cycle (Day 12)Time (h) N-NH4 N-NO2 N-NO3

(mg/L) (mg/L) (mg/L)

0 12.5 0.0 0.0

0.2 7.8 0.1 0.8

1.3 5.1 0.3 2.7

2.3 2.0 0.6 4.0

3.3 0.3 0.2 6.7

4.3 0.2 0.0 7.7

5.3 0.2 0.0 7.6

6.3 0.2 0.0 7.9

7.6 0.2 2.8 2.0

7.8 0.2 0.0 0.0

8.3 0.2 0.0 0.0

9.3 0.2 0.0 0.0

10.3 0.2 0.0 0.0

12.0 0.2 0.0 0.0

y = -3.2385x + 10.115R² = 0.8862

0.0

2.0

4.0

6.0

8.0

10.0

12.0

14.0

0 2 4 6 8 10 12

N-NH4 (mg/L)

N-NO2 (mg/L)

N-NO3 (mg/L)

Linear (N-NH4)

Time (h)

Co

nce

ntr

atio

n (

mg/

L)

5 March within a cycle (Day 16)Time (h) N-NH4 N-NO2 N-NO3

(mg/L) (mg/L) (mg/L)

0 12.6871 0.0 0.0

0.2 10.0 0.0 0.8

1.3 6.2 0.0 1.6

2.3 4.3 0.0 3.6

2.8 3.0 0.2 3.8

3.3 1.9 0.3 4.9

3.6 1.0 0.2 5.4

3.8 0.7 0.2 5.9

4.1 0.4 0.0 6.6

4.3 0.3 0.3 7.2

4.6 0.0 0.0 7.4

5.3 0.0 0.0 7.4

6.3 0.0 0.0 7.9

7.6 0.0 1.2 4.3

8.3 0 0 0

9.3 0 0 0

10.3 0 0 0

11.3 0 0 0

y = -2.6876x + 10.921R² = 0.9613

0

2

4

6

8

10

12

14

0 2 4 6 8 10 12

N-NH4 (mg/L)

N-NO2 (mg/L)

N-NO3 (mg/L)

Linear (N-NH4)

Time (h)

Co

nce

ntr

atio

n (

mg/

L)

Figure A.5: Ammonium conversion rate calculation in period 2 day 68

Figure A.6: Ammonium conversion rate calculation in period 2 day 75

26 April within a cycle (Day 68)Time (h) N-NH4 N-NO2 N-NO3

(mg/L) (mg/L) (mg/L)

0.0 7.8 0.0 3.9

0.3 4.6 0.0 0.6

0.7 4.5 0.0 0.0

1.3 3.0 0.0 0.8

1.8 1.5 0.0 1.4

2.3 0.0 0.0 2.2

3.3 0.0 0.0 3.1

4.3 0.0 0.0 3.2

5.0 5.8 0.0 3.2

5.3 2.6 0.0 0.4

5.7 2.6 0.0 0.4

6.3 1.3 0.0 0.9

6.8 0.0 0.0 1.7

7.3 0.0 0.0 2.4

8.3 0.0 0.0 2.2

9.3 0.0 0.0 2.2

10.3 0.0 0.0 2.3

11.3 0.0 0.0 2.2

y = -2.9433x + 19.497R² = 0.7471

y = -2.9564x + 6.7233R² = 0.8862

0.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

9.0

0 2 4 6 8 10 12

N-NH4 (mg/L)

N-NO2 (mg/L)

N-NO3 (mg/L)

Linear (ka)

Linear (kb)

Time (h)

Co

nce

ntr

atio

n (

mg/

L)

3 May within a cycle (Day 75)Time (h) N-NH4 N-NO2 N-NO3

(mg/L) (mg/L) (mg/L)

0.0 8.1 0.0 1.8

0.3 4.5 0.0 0.0

0.9 4.4 0.0 0.5

1.3 2.9 0.0 0.8

1.8 1.4 0.0 1.7

2.8 0.0 0.0 3.7

3.8 0.0 0.0 3.8

5.0 6.1 0.0 3.9

5.3 2.8 1.2 0.0

5.9 2.1 0.0 0.7

6.3 1.0 0.0 1.1

6.8 0.0 0.0 1.9

7.8 0.0 0.0 2.4

8.8 0.0 0.0 2.5

10.0 0.0 0.0 2.5

11.0 0.0 0.0 2.6

y = -3.3755x + 22.023R² = 0.8249

y = -3.2618x + 7.0282R² = 0.8614

0.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

9.0

0 2 4 6 8 10 12

N-NH4 (mg/L)

N-NO2 (mg/L)

N-NO3 (mg/L)

Linear (ka)

Linear (kb)

Time (h)

Co

nce

ntr

atio

n (

mg/

L)

Figure A.7: Ammonium conversion rate calculation in period 2 day 89

Figure A.8: Ammonium conversion rate calculation in period 3 day 110

17 May within a cycle (Day 89)Time (h) N-NH4 N-NO2 N-NO3

(mg/L) (mg/L) (mg/L)

0.0 6.9 0.0 1.6

0.3 4.1 0.3 0.8

0.9 3.2 0.1 0.9

1.3 2.5 0.1 1.3

1.8 1.4 0.1 2.0

2.8 0.0 0.0 3.7

3.8 0.0 0.0 4.1

5.0 5.1 0.0 4.4

5.3 3.0 0.8 2.4

5.9 2.4 1.0 1.4

6.3 1.5 0.2 2.8

6.8 0.4 0.1 4.0

7.8 0.0 0.0 4.5

8.8 0.0 0.0 4.6

10.0 0.0 0.0 4.5

11.0 0.0 0.0 4.8

11.5 0.0 0.0 4.6

y = -2.3807x + 16.414R² = 0.9191

y = -2.7456x + 5.951R² = 0.8891

0.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

0 2 4 6 8 10 12

N-NH4 (mg/L)

N-NO2 (mg/L)

N-NO3 (mg/L)

Linear (ka)

Linear (kb)

Time (h)

Co

nce

ntr

atio

n (

mg/

L)

7 June within a cycle (Day 110)Time (h) N-NH4 N-NO2 N-NO3

(mg/L) (mg/L) (mg/L)

0.0 7.6 0.0 1.5

0.3 5.2 0.1 0.1

0.9 5.1 0.0 0.0

1.3 3.8 0.1 0.0

1.8 3.7 0.0 0.4

2.3 0.0 1.8 1.6

3.0 0.0 1.2 2.3

5.0 5.8 0.1 3.3

5.3 4.2 0.7 1.6

5.9 4.1 0.0 0.0

6.3 3.7 0.0 0.1

6.8 2.2 0.6 0.5

7.8 0.0 1.7 1.4

8.8 0.0 0.5 2.3

10.0 0.0 0.0 3.0

11.0 0.0 0.0 3.2

12.0 0.0 0.0 3.1

y = -1.6926x + 13.939R² = 0.8484

y = -2.0065x + 6.8129R² = 0.8128

0.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

9.0

0 2 4 6 8 10 12

N-NH4 (mg/L)

N-NO2 (mg/L)

N-NO3 (mg/L)

Linear (ka)

Linear (kb)

Time (h)

Co

nce

ntr

atio

n (

mg/

L)

Figure A.9: Ammonium conversion rate calculation in period 3 day 117

Figure A.10: Ammonium conversion rate calculation in period 3 day 123

14 June within a cycle (Day 117)Time (h) N-NH4 N-NO2 N-NO3

(mg/L) (mg/L) (mg/L)

0.0 7.9 0.0 2.0

0.3 6.3 0.4 0.1

0.9 4.3 0.2 0.0

1.3 3.7 0.3 0.5

1.8 3.0 0.9 1.0

2.3 1.1 1.6 1.6

2.8 0.5 1.6 1.8

3.3 0.4 0.9 2.5

4.6 0.4 0.1 3.3

5.0 6.1 0.0 3.3

5.3 5.1 0.0 1.6

5.9 5.2 0.8 0.8

6.3 3.6 0.2 0.2

6.9 3.2 0.5 1.3

7.3 1.9 1.0 1.8

7.8 0.3 1.1 2.5

8.3 0.0 0.8 2.9

10.0 0.0 0.3 3.0

11.0 0.0 0.0 3.2

12.0 0.0 0.0 2.0

y = -1.9342x + 15.939R² = 0.9339

y = -2.6118x + 7.3515R² = 0.9775

0.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

9.0

0 2 4 6 8 10 12

N-NH4 (mg/L)

N-NO2 (mg/L)

N-NO3 (mg/L)

Linear (ka)

Linear (kb)

Time (h)

Co

nce

ntr

atio

n (

mg/

L)

20 June within a cycle (Day 123)Time (h) N-NH4 N-NO2 N-NO3

(mg/L) (mg/L) (mg/L)

0.0 7.6 0.0 1.3

0.3 6.6 0.4 0.7

0.9 4.2 0.1 0.4

1.3 4.1 0.3 0.9

1.8 2.0 0.4 1.3

2.3 1.9 0.3 1.7

2.8 1.2 0.4 2.3

3.3 1.1 0.4 2.7

4.3 0.0 0.1 3.4

5.0 5.9 0.1 3.5

5.3 4.6 0.3 1.5

5.9 4.3 0.5 0.1

6.8 2.8 0.1 1.0

7.3 1.5 0.2 1.3

7.8 1.3 0.2 1.7

8.8 0.4 0.3 2.4

10.0 0.0 0.2 2.8

11.0 0.0 0.1 2.7

11.5 0.0 0.1 2.5

y = -1.4228x + 12.543R² = 0.9582

y = -2.2912x + 7.0915R² = 0.9412

0.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

9.0

0 2 4 6 8 10 12

N-NH4 (mg/L)

N-NO2 (mg/L)

N-NO3 (mg/L)

Linear (ka)

Linear (kb)

Time (h)

Co

nce

ntr

atio

n (

mg/

L)

Figure A.11: Ammonium conversion rate calculation in period 3 day 131

Figure A.12: Ammonium conversion rate calculation in period 3 day 138

28 June within a cycle (Day 131)Time (h) N-NH4 N-NO2 N-NO3

(mg/L) (mg/L) (mg/L)

0.0 8.3 0.0 1.7

0.3 7.2 0.3 0.8

0.9 6.5 0.0 0.0

1.3 5.5 0.0 0.3

1.8 3.2 0.1 1.3

2.3 3.0 0.1 2.0

2.8 1.2 0.2 2.5

3.3 0.9 0.1 3.4

5.0 6.9 0.0 3.3

5.3 5.2 0.3 1.9

5.9 5.1 0.0 0.1

6.3 4.3 0.0 0.6

6.8 2.5 0.1 0.8

7.3 1.5 0.0 1.5

7.8 0.6 0.2 2.5

8.3 0.6 0.0 2.9

10.0 0.4 0.0 2.7

11.0 0.7 0.0 2.8

11.5 0.0 0.0 2.6

y = -1.9464x + 16.225R² = 0.9534

y = -2.3486x + 8.2284R² = 0.964

0.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

9.0

0 2 4 6 8 10 12

N-NH4 (mg/L)

N-NO2 (mg/L)

N-NO3 (mg/L)

Linear (ka)

Linear (kb)

Time (h)

Co

nce

ntr

atio

n (

mg/

L)

5 July within a cycle (Day 138)Time (h) N-NH4 N-NO2 N-NO3

(mg/L) (mg/L) (mg/L)

0.0 7.9 0.0 1.5

0.3 5.7 0.3 1.0

0.9 5.2 0.0 0.1

1.3 4.5 0.0 0.4

1.8 2.7 0.1 1.3

2.3 1.8 0.1 1.7

2.8 1.0 0.2 2.8

3.3 0.3 0.1 3.6

5.0 6.2 0.0 3.4

5.3 4.3 0.3 1.6

5.9 4.1 0.0 0.3

6.3 2.9 0.0 0.5

6.8 2.9 0.1 1.1

7.3 0.8 0.0 1.9

7.8 0.0 0.2 2.6

8.3 0.0 0.0 3.0

10.0 0.5 0.0 2.6

11.0 0.0 0.0 2.3

12.0 0.1 0.0 2.6

y = -1.929x + 15.334R² = 0.901

y = -2.1815x + 7.1548R² = 0.9666

0.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

9.0

0 2 4 6 8 10 12

N-NH4 (mg/L)

N-NO2 (mg/L)

N-NO3 (mg/L)

Linear (ka)

Linear (kb)

Time (h)

Co

nce

ntr

atio

n (

mg/

L)

Figure A.13: Ammonium conversion rate calculation in period 4 day 151

Figure A.14: Ammonium conversion rate calculation in period 4 day 165

18 July within a cycle (Day 151)Time (h) N-NH4 N-NO2 N-NO3

(mg/L) (mg/L) (mg/L)

0.0 6.3 0.0 0.9

0.3 3.7 0.3 0.6

0.9 3.0 0.2 0.2

1.3 2.3 0.4 0.4

1.8 1.6 0.5 0.7

2.3 0.2 0.5 0.9

2.8 0.0 0.8 1.6

3.3 0.0 0.6 5.7

5.0 4.7 0.2 1.8

5.3 3.1 0.4 1.0

5.9 2.6 0.2 0.1

6.3 1.5 0.4 0.2

6.8 0.9 0.5 0.6

7.3 0.0 0.6 1.0

7.8 0.0 0.8 1.3

8.3 0.0 0.6 1.7

10.0 0.0 0.2 1.5

11.0 0.0 0.2 1.1

12.0 0.0 0.3 1.5

y = -1.9358x + 13.965R² = 0.9331

y = -2.2365x + 5.3752R² = 0.9109

0.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

0 2 4 6 8 10 12

N-NH4 (mg/L)

N-NO2 (mg/L)

N-NO3 (mg/L)

Linear (ka)

Linear (kb)

Time (h)

Co

nce

ntr

atio

n (

mg/

L)

1 August within a cycle (Day165)Time (h) N-NH4 N-NO2 N-NO3

(mg/L) (mg/L) (mg/L)

0.0 7.1 0.0 0.3

0.3 3.7 2.9 0.0

0.9 3.0 0.3 0.0

1.3 2.3 0.7 0.0

1.8 1.6 1.7 0.2

2.3 0.2 2.8 0.2

2.8 0.0 3.2 0.4

3.3 0.0 4.7 0.5

5.0 5.3 5.0 0.9

5.3 3.1 2.7 0.0

5.9 2.6 1.0 0.2

6.3 1.5 1.2 1.9

7.3 0.9 2.6 0.3

8.3 0.0 4.4 1.0

10.0 0.0 4.1 1.4

11.0 0.0 3.9 0.6

12.0 0.0 3.9 0.8

y = -1.6835x + 12.763R² = 0.8208

y = -2.4549x + 5.7484R² = 0.8671

0.0

1.0

2.0

3.0

4.0

5.0

6.0

7.0

8.0

0 2 4 6 8 10 12

N-NH4 (mg/L)

N-NO2 (mg/L)

N-NO3 (mg/L)

Linear (ka)

Linear (kb)

Time (h)

Co

nce

ntr

atio

n (

mg/

L)

Appendix B

B.1 R-output for U-test and H-test for Ammonium removal rate

> # Any difference in ammonium removal rate between period1 and period2? > > ############# > # Read file # > data<-read.table("AmmonremovrateUtest.txt",header=TRUE) > period1<-data$rate[data$period=="period1"] > period2<-data$rate[data$period=="period2"] > > ############################################################# > # Test for normal distribution (beware: small sample size!) # > shapiro.test(period1) Shapiro-Wilk normality test data: period1 W = 0.9631, p-value = 0.7982 > shapiro.test(period2) Shapiro-Wilk normality test data: period2 W = 0.8663, p-value = 0.2117 > > #################################### > # QQ-plots, histograms and boxplot # > layout(matrix(c(1,2,3,4,5,6),3,2,byrow = TRUE)) > qqnorm(period1,main="Normal QQ-plot of period1");qqline(period1) > qqnorm(period2,main="Normal QQ-plot of period2");qqline(period2) > hist(period1) > hist(period2) > boxplot(rate~period,data=data) > > ####################### > # Mann-Whitney U-test # > wilcox.test(rate~period,data=data) Wilcoxon rank sum test with continuity correction data: rate by period W = 24, p-value = 0.01278 alternative hypothesis: true location shift is not equal to 0 Warning message: In wilcox.test.default(x = c(2.3, 2.5, 2.6, 2.1), y = c(1.2, 1.1, : cannot compute exact p-value with ties

# Kruskal-Wallis H-test (stacked arramgement) > rm(list=ls(all.names=TRUE)) > > # Do the three period differ with respect to spec amon rate? > > ############# > # Read file # > data<-read.table("AmmonremovrateHtest.txt",header=TRUE) > > ################################################ > # Make a factor out of the grouping variables! # > data$period<-factor(data$period) > > period2<-data$rate[data$period=="period2"] > period3<-data$rate[data$period=="period3"] > period4<-data$rate[data$period=="period4"] > > ############ > # Boxplots # > boxplot(rate~period,data=data) > > ######################### > # Kruskal-Wallis H-test # > kruskal.test(rate~period,data=data) Kruskal-Wallis rank sum test data: rate by period Kruskal-Wallis chi-squared = 13.5316, df = 2, p-value = 0.001153 >

> ############################ > # Post-hoc test from ANOVA # > fit<-aov(rate~period,data=data) > TukeyHSD(fit) Tukey multiple comparisons of means 95% family-wise confidence level Fit: aov(formula = rate ~ period, data = data) $period diff lwr upr p adj period3-period2 1.083333 -0.8125504 2.979217 0.3313478 period4-period2 3.533333 1.1634786 5.903188 0.0036751 period4-period3 2.450000 0.2779923 4.622008 0.0259380

B.2 R-output for U-test and H-test for VSS

> # Mann-Whitney U-test for independent samples (stacked data) > rm(list=ls(all.names=TRUE)) > > # Any difference in VSS between period1 and period2? > > ############# > # Read file # > data<-read.table("VSS_Utest.txt",header=TRUE) > period1<-data$vss[data$period=="period1"] > period2<-data$vss[data$period=="period2"] > > ############################################################# > # Test for normal distribution (beware: small sample size!) # > shapiro.test(period1) Shapiro-Wilk normality test data: period1 W = 0.8567, p-value = 0.0001314 > shapiro.test(period2) Shapiro-Wilk normality test data: period2 W = 0.9568, p-value = 0.3772 > > #################################### > # QQ-plots, histograms and boxplot # > layout(matrix(c(1,2,3,4,5,6),3,2,byrow = TRUE)) > qqnorm(period1,main="Normal QQ-plot of period1");qqline(period1) > qqnorm(period2,main="Normal QQ-plot of period2");qqline(period2) > hist(period1) > hist(period2) > boxplot(vss~period,data=data) > > ####################### > # Mann-Whitney U-test # > wilcox.test(vss~period,data=data) Wilcoxon rank sum test with continuity correction data: vss by period W = 228.5, p-value = 0.0004823 alternative hypothesis: true location shift is not equal to 0 Warning message: In wilcox.test.default(x = c(1.1, 1.2, 1, 1.1, 1.1, 0.9, 1.1, 1.2, : cannot compute exact p-value with ties

######################### > # Kruskal-Wallis H-test # > kruskal.test(VSS~period,data=data) Kruskal-Wallis rank sum test data: VSS by period Kruskal-Wallis chi-squared = 41.1926, df = 2, p-value = 1.135e-09 > > ############################ > # Post-hoc test from ANOVA # > fit<-aov(VSS~period,data=data) > TukeyHSD(fit) Tukey multiple comparisons of means 95% family-wise confidence level Fit: aov(formula = VSS ~ period, data = data) $period diff lwr upr p adj period3-period2 -1.36133333 -1.6787313 -1.0439354 0.0000000 period4-period2 -1.37948718 -1.6938813 -1.0650931 0.0000000 period4-period3 -0.01815385 -0.3292611 0.2929534 0.9893084

B.3 R-output for U-test and H-test for Chlorophyll

> # Mann-Whitney U-test for independent samples (stacked data) > rm(list=ls(all.names=TRUE)) > > # Any difference in VSS between period1 and period2? > > ############# > # Read file # > data<-read.table("chloro-Utest.txt",header=TRUE) > period1<-data$chlorophyll[data$period=="period1"] > period2<-data$chlorophyll[data$period=="period2"] > > ############################################################# > # Test for normal distribution (beware: small sample size!) # > shapiro.test(period1) Shapiro-Wilk normality test data: period1 W = 0.8978, p-value = 0.00297 > shapiro.test(period2) Shapiro-Wilk normality test data: period2 W = 0.926, p-value = 0.1013 > > #################################### > # QQ-plots, histograms and boxplot # > layout(matrix(c(1,2,3,4,5,6),3,2,byrow = TRUE)) > qqnorm(period1,main="Normal QQ-plot of period1");qqline(period1) > qqnorm(period2,main="Normal QQ-plot of period2");qqline(period2) > hist(period1) > hist(period2) > boxplot(chlorophyll~period,data=data) > > ####################### > # Mann-Whitney U-test # > wilcox.test(chlorophyll~period,data=data) Wilcoxon rank sum test with continuity correction data: chlorophyll by period W = 210.5, p-value = 0.003026 alternative hypothesis: true location shift is not equal to 0 Warning message: In wilcox.test.default(x = c(5.2, 8.7, 11.3, 9, 10.8, 11.8, 10.3, : cannot compute exact p-value with ties

> ######################### > # Kruskal-Wallis H-test # > kruskal.test(chlorophyll~period,data=data) Kruskal-Wallis rank sum test data: chlorophyll by period Kruskal-Wallis chi-squared = 45.0964, df = 2, p-value = 1.612e-10 > > ############################ > # Post-hoc test from ANOVA # > fit<-aov(chlorophyll~period,data=data) > TukeyHSD(fit) Tukey multiple comparisons of means 95% family-wise confidence level Fit: aov(formula = chlorophyll ~ period, data = data) $period diff lwr upr p adj period3-period2 -20.183712 -24.230904 -16.136520 0.0000000 period4-period2 -22.967045 -27.014238 -18.919853 0.0000000 period4-period3 -2.783333 -6.741565 1.174899 0.2182491

B.4 R-output for U-test and H-test for Oxygen Production

> # Mann-Whitney U-test for independent samples (stacked data) > rm(list=ls(all.names=TRUE)) > > # Any difference in Oxprod between period1 and period2? > > ############# > # Read file # > data<-read.table("Oxprod-Utest.txt",header=TRUE) > period1<-data$oxygen[data$period=="period1"] > period2<-data$oxygen[data$period=="period2"] > > ############################################################# > # Test for normal distribution (beware: small sample size!) # > shapiro.test(period1) Shapiro-Wilk normality test data: period1 W = 0.8411, p-value = 0.07739 > shapiro.test(period2) Shapiro-Wilk normality test data: period2 W = 0.936, p-value = 0.3025 > > #################################### > # QQ-plots, histograms and boxplot # > layout(matrix(c(1,2,3,4,5,6),3,2,byrow = TRUE)) > qqnorm(period1,main="Normal QQ-plot of period1");qqline(period1) > qqnorm(period2,main="Normal QQ-plot of period2");qqline(period2) > hist(period1) > hist(period2) > boxplot(oxygen~period,data=data) > > ####################### > # Mann-Whitney U-test # > wilcox.test(oxygen~period,data=data) Wilcoxon rank sum test with continuity correction data: oxygen by period W = 128, p-value = 9.737e-05 alternative hypothesis: true location shift is not equal to 0 Warning message: In wilcox.test.default(x = c(4.3, 4.1, 3.7, 4.2, 4, 4.3, 3.7, 4.2 : cannot compute exact p-value with ties

> ######################### > # Kruskal-Wallis H-test # > kruskal.test(oxygen~period,data=data) Kruskal-Wallis rank sum test data: oxygen by period Kruskal-Wallis chi-squared = 31.6765, df = 2, p-value = 1.323e-07 > > ############################ > # Post-hoc test from ANOVA # > fit<-aov(oxygen~period,data=data) > TukeyHSD(fit) Tukey multiple comparisons of means 95% family-wise confidence level Fit: aov(formula = oxygen ~ period, data = data) $period diff lwr upr p adj period3-period2 4.98750 -1.329567 11.30457 0.1459520 period4-period2 16.23333 9.410121 23.05655 0.0000026 period4-period3 11.24583 4.422621 18.06905 0.0007246

B.4 R-output for Normal Distribution test and H-test for Nitrogen removal

# Read file # > data<-read.table("nitrogenremoval.txt",header=TRUE)

> # Read file # > data<-read.table("nitrogenremoval.txt",header=TRUE) > period2<-data$efficiency[data$period=="period2"] > period3<-data$efficiency[data$period=="period3"] > period4<-data$efficiency[data$period=="period4"] > # Test for normal distribution (beware: small sample size!) # > shapiro.test(period2) Shapiro-Wilk normality test data: period2 W = 0.8185, p-value = 0.00047 > shapiro.test(period3) Shapiro-Wilk normality test data: period3 W = 0.9545, p-value = 0.4137 > shapiro.test(period4) Shapiro-Wilk normality test data: period4 W = 0.9051, p-value = 0.1573 > # QQ-plots, histograms and boxplot # > layout(matrix(c(1,2,3,4,5,6),3,2,byrow = TRUE)) > qqnorm(period2,main="Normal QQ-plot of period2");qqline(period2) > qqnorm(period3,main="Normal QQ-plot of period3");qqline(period3) > qqnorm(period4,main="Normal QQ-plot of period4");qqline(period4) > hist(period2) > hist(period3) > hist(period4) > boxplot(efficiency~period,data=data)

> # Kruskal-Wallis H-test (stacked arramgement) > rm(list=ls(all.names=TRUE)) > > # Do the three period differ with respect to nitrogen removal? > > ############# > # Read file # > data<-read.table("nitrogenremoval.txt",header=TRUE) > > ################################################ > # Make a factor out of the grouping variables! # > data$period<-factor(data$period) > > period2<-data$efficiency[data$period=="period2"] > period3<-data$efficiency[data$period=="period3"] > period4<-data$efficiency[data$period=="period4"] > > ############ > # Boxplots # > boxplot(efficiency~period,data=data) > > ######################### > # Kruskal-Wallis H-test # > kruskal.test(efficiency~period,data=data) Kruskal-Wallis rank sum test data: efficiency by period Kruskal-Wallis chi-squared = 9.4334, df = 2, p-value = 0.008945 > > ############################

> # Post-hoc test from ANOVA # > fit<-aov(efficiency~period,data=data) > TukeyHSD(fit) Tukey multiple comparisons of means 95% family-wise confidence level Fit: aov(formula = efficiency ~ period, data = data) $period diff lwr upr p adj period3-period2 15.575810 3.870852 27.28077 0.0062447 period4-period2 17.775077 4.253707 31.29645 0.0069751 period4-period3 2.199267 -11.755696 16.15423 0.9238233

Appendix C

C.1 Daily TSS/VSS concentration

Date Day TSS (g/l) VSS (g/l)

18/Feb/13 1 3.14 1.14

19/Feb/13 2 2.80 1.18

20/Feb/13 3 2.60 1.05

21/Feb/13 4 2.53 1.09

22/Feb/13 5 2.03 1.05

25/Feb/13 8 1.84 0.95

26/Feb/13 9 1.96 1.08

27/Feb/13 10 2.24 1.24

28/Feb/13 11 2.48 1.32

1/Mar/13 12 2.18 1.25

4/Mar/13 15 1.90 1.22

5/Mar/13 16 2.06 1.27

6/Mar/13 17 1.97 1.26

7/Mar/13 18 1.87 1.22

11/Mar/13 22 1.41 1.00

12/Mar/13 23 1.88 1.26

13/Mar/13 24 1.65 1.17

14/Mar/13 25 1.63 1.14

15/Mar/13 26 1.63 1.14

18/Mar/13 29 1.50 1.15

19/Mar/13 30 1.56 1.17

20/Mar/13 31 2.26 1.66

21/Mar/13 32 1.69 1.41

25/Mar/13 36 1.08 0.82

28/Mar/13 39 3.47 2.21

2/Apr/13 44 3.78 2.51

3/Apr/13 45 3.60 2.39

4/Apr/13 46 4.15 2.70

5/Apr/13 47 3.52 2.33

8/Apr/13 50 3.82 2.58

9/Apr/13 51 3.46 2.40

10/Apr/13 52 4.05 2.79

11/Apr/13 53 3.51 2.50

12/Apr/13 54 3.50 2.49

15/Apr/13 57 2.87 2.16

16/Apr/13 58 2.81 2.15

17/Apr/13 59 4.23 3.15

18/Apr/13 60 3.39 2.55

19/Apr/13 61 3.01 2.28

22/Apr/13 64 2.21 1.75

23/Apr/13 65 3.08 2.32

24/Apr/13 66 3.15 2.45

25/Apr/13 67 3.36 2.64

26/Apr/13 68 3.16 2.49

29/Apr/13 71 2.40 1.95

1/May/13 73 3.85 3.19

2/May/13 74 3.48 2.94

3/May/13 75 3.73 2.99

6/May/13 78 2.78 2.33

7/May/13 79 4.16 3.36

8/May/13 80 3.49 2.82

10/May/13 82 3.15 2.54

13/May/13 85 2.14 1.77

14/May/13 86 3.42 2.76

15/May/13 87 3.20 2.63

16/May/13 88 2.72 2.24

17/May/13 89 2.90 2.44

21/May/13 93 1.75 1.48

22/May/13 94 2.97 2.48

23/May/13 95 2.84 2.40

24/May/13 96 2.70 2.21

27/May/13 99 1.67 1.43

28/May/13 100 1.06 0.95

3/Jun/13 106 1.50 1.25

4/Jun/13 107 1.30 1.09

5/Jun/13 108 1.23 1.03

6/Jun/13 109 1.17 1.07

7/Jun/13 110 1.72 1.46

10/Jun/13 113 1.26 1.11

11/Jun/13 114 1.33 1.19

12/Jun/13 115 0.68 0.64

13/Jun/13 116 0.76 0.64

14/Jun/13 117 1.45 1.27

17/Jun/13 120 0.52 0.46

18/Jun/13 121 1.37 1.23

19/Jun/13 122 0.87 0.79

20/Jun/13 123 0.52 0.51

21/Jun/13 124 1.08 0.96

24/Jun/13 127 0.71 0.66

25/Jun/13 128 1.53 1.35

26/Jun/13 129 1.34 1.19

27/Jun/13 130 1.48 1.29

28/Jun/13 131 1.61 1.44

1/Jul/13 134 0.56 0.50

2/Jul/13 135 1.19 1.06

3/Jul/13 136 0.91 0.83

4/Jul/13 137 1.02 0.92

5/Jul/13 138 1.06 0.98

8/Jul/13 141 0.54 0.54

9/Jul/13 142 1.53 1.39

10/Jul/13 143 1.05 0.96

11/Jul/13 144 0.79 0.72

12/Jul/13 145 0.53 0.50

15/Jul/13 148 0.60 0.56

16/Jul/13 149 1.39 1.23

17/Jul/13 150 0.82 0.73

18/Jul/13 151 0.33 0.29

19/Jul/13 152 0.26 0.24

22/Jul/13 155 0.61 0.56

23/Jul/13 156 0.24 0.22

24/Jul/13 157 0.24 0.22

25/Jul/13 158 1.67 1.48

26/Jul/13 159 1.58 1.44

29/Jul/13 162 1.28 1.14

30/Jul/13 163 1.78 1.62

31/Jul/13 164 1.48 1.34

1/Aug/13 165 1.071 0.969

5/Aug/13 169 0.826 0.754

6/Aug/13 170 1.296 1.106

7/Aug/13 171 1.627 1.426

8/Aug/13 172 1.603 1.42

9/Aug/13 173 1.589 1.488

12/Aug/13 176 1.034 1.016

13/Aug/13 177 1.37 1.234

14/Aug/13 178 1.007 0.898

C.2 Daily Chlorophyll-a concentration

Date Day Chlorophyll

mg/l 18-Feb 1 5.15

20-Feb 3 8.75

21-Feb 4 11.27

22-Feb 5 9.03

25-Feb 8 10.83

26-Feb 9 11.81

27-Feb 10 10.31

28-Feb 11 12.21

1-Mar 12 9.70

4-Mar 15 10.33

5-Mar 16 13.07

6-Mar 17 11.77

7-Mar 18 9.99

11-Mar 22 15.14

12-Mar 23 16.43

13-Mar 24 15.27

18/03/13 29 6.44

19/03/13 30 9.18

20/03/13 31 12.95

21/03/13 32 13.10

28/03/13 39 21.46

2/04/13 44 33.15

3/04/13 45 28.79

4/04/13 46 32.19

5/04/13 47 30.19

8/04/13 50 24.79

9/04/13 51 26.49

10/04/13 52 37.30

11/04/13 53 29.53

12/04/13 54 33.08

15/04/13 57 32.69

16/04/13 58 27.01

17/04/13 59 42.77

18/04/13 60 30.78

19/04/13 61 29.30

22/04/13 64 24.79

23/04/13 65 37.44

24/04/13 66 35.00

25/04/13 67 35.45

26/04/13 68 27.23

29/04/13 71 22.13

1/05/13 73 29.60

2/05/13 74 30.34

3/05/13 75 30.56

6/05/13 78 20.94

7/05/13 79 39.07

8/05/13 80 35.74

10/05/13 82 31.30

13/05/13 85 20.28

14/05/13 86 28.64

15/05/13 87 23.75

16/05/13 88 22.35

17/05/13 89 20.57

21/05/13 93 10.29

22/05/13 94 35.37

23/05/13 95 36.56

24/05/13 96 33.15

28/05/13 100 11.54

4/06/13 107 9.55

5/06/13 108 10.43

6/06/13 109 9.77

10/06/13 113 10.66

11/06/13 114 10.88

12/06/13 115 4.07

13/06/13 116 5.85

14/06/13 117 9.40

17/06/13 120 3.03

18/06/13 121 12.14

19/06/13 122 5.77

20/06/13 123 3.77

21/06/13 124 8.88

24/06/13 127 6.07

25/06/13 128 11.99

26/06/13 129 9.62

27/06/13 130 10.66

28/06/13 131 10.88

1/07/13 134 2.44

2/07/13 135 9.10

3/07/13 136 7.10

4/07/13 137 5.85

5/07/13 138 7.33

8/07/13 141 3.70

9/07/13 142 5.11

10/07/13 143 2.66

11/07/13 144 1.48

12/07/13 145 0.96

15/07/13 148 3.034

16/07/13 149 8.51

17/07/13 150 4.81

18/07/13 151 1.48

19/07/13 152 1.184

23/07/13 156 0.148

24/07/13 157 0.37

25/07/13 158 1.11

26/07/13 159 1.258

29/07/13 162 0.814

30/07/13 163 2.96

31/07/13 164 3.478

1/08/13 165 3.478

5/08/13 169 2.368

6/08/13 170 8.362

7/08/13 171 16.576

8/08/13 172 12.062

9/08/13 173 13.246

12/08/13 176 15.54

13/08/13 177 10.878

C.2 Daily Influent and Effluent COD (mg/L)

Day Effluent Influent

21/03/2013 52.0

22/03/2013 52.3

25/03/2013 54.5

26/03/2013 12.9

27/03/2013 29.8

28/03/2013 0.0

02/04/2013 64.8

03/04/2013 44.0

04/04/2013 56.0

05/04/2013 37.9

08/04/2013 23.2

09/04/2013 40.4

10/04/2013 29.3

11/04/2013 21.7

12/04/2013 27.3

15/04/2013 0.0 217.3

16/04/2013 4.2 270.0

17/04/2013 8.7 199.3

18/04/2013 22.3 219.7

19/04/2013 19.2 195.3

22/04/2013 19.5 138.0

23/04/2013 13.2 150.0

24/04/2013 18.2 146.7

25/04/2013 17.9 164.0

26/04/2013 12.0 130.7

29/04/2013 19.8 184.0

01/05/2013 22.3 190.7

02/05/2013 27.9 177.3

03/05/2013 20.1 198.0

06/05/2013 0.0 197.3

07/05/2013 7.0 172.0

10/05/2013 13.8 172.7

13/05/2013 22.2 206.7

14/05/2013 48.7 178.0

15/05/2013 25.2 186.0

16/05/2013 12.2 209.3

17/05/2013 45.2 185.3

21/05/2013 39.9 210.7

22/05/2013 21.1 197.3

23/05/2013 26.9 246.0

24/05/2013 29.9 189.3

27/05/2013 35.8 165.3

28/05/2013 0.0 160.7

04/06/2013 30.5 192.7

05/06/2013 0.0 164.0

06/06/2013 30.5 162.0

07/06/2013 12.2 176.7

10/06/2013 0.0 182.7

11/06/2013 0.0 194.7

12/06/2013 0.0 159.3

13/06/2013 0.0 174.7

14/06/2013 21.1 192.0

17/06/2013 76.2 197.3

18/06/2013 47.1 195.3

19/06/2013 22.4 185.3

21/06/2013 43.3 188.7

24/06/2013 31.4 234.7

25/06/2013 40.0 198.7

26/06/2013 41.0 272.7

27/06/2013 41.4 174.0

28/06/2013 33.8 272.7

01/07/2013 44.8 218.0

02/07/2013 29.0 176.0

03/07/2013 33.3 172.0

04/07/2013 75.7 178.0

05/07/2013 41.4 170.7

08/07/2013 17.6 171.3

09/07/2013 14.3 174.7

10/07/2013 13.8 168.0

11/07/2013 9.0 170.0

15/07/2013 40.5 174.7

16/07/2013 0.0 193.3

17/07/2013 0.0 196.7

18/07/2013 1.0 188.0

19/07/2013 0.0 222.7

22/07/2013 2.4 184.0

23/07/2013 19.0 190.0

24/07/2013 77.1 184.7

29/07/2013 15.5 129.3

30/07/2013 7.6 173.3

31/07/2013 4.0 179.0

01/08/2013 32.1 182.7

05/08/2013 14.8 171.7

06/08/2013 18.3 189.7

07/08/2013 21.0 177.7

08/08/2013 7.9 176.0

09/08/2013 26.4 129.0

Appendix D

Oxygen production calculation

C(t) = A*(1-exp (-B*t))+Co*exp(-B*t)

A B A B1st

aerobic

2nd

aerobicavg Std dev

1st

aerobic

2nd

aerobicavg Std dev

1st

aerobic

2nd

aerobic

1st

aerobic

2nd

aerobicavg Std dev

1st

aerobic

2nd

aerobicavg Std dev A B A B

1st

aerobic

2nd

aerobicavg Std dev

1st

aerobic

2nd

aerobicavg Std dev

4/04/13 46 2.7 18.0 1.1 11.6 4.3 3.7 15.3 5.6 -57.3 0.6 36.1 13.4

5/04/13 47 2.3 20.5 0.7 9.4 4.1 3.7 13.1 5.7 -29.6 1.1 36.7 15.8

8/04/13 50 2.6 19.3 0.8 9.5 3.7 3.7 13.2 5.1 -17.1 1.7 39.3 15.3

9/04/13 51 2.4 18.0 1.0 10.1 4.2 3.7 13.8 5.7 -30.5 1.0 35.5 14.8

12/04/13 54 2.5 16.2 1.2 9.9 4.0 3.7 13.6 5.5 -12.8 2.6 49.4 19.8

15/04/13 57 2.2 16.0 1.1 9.3 4.3 3.7 13.0 6.0 -20.2 1.9 48.1 22.3

16/04/13 58 2.2 15.1 1.1 8.0 3.7 3.7 11.7 5.5 -36.0 1.0 39.5 18.3

19/04/13 61 2.3 16.6 1.1 9.7 4.2 3.7 13.4 5.9 -38.1 0.9 35.9 15.7

26/04/13 68 2.5 15.1 0.8 15.4 1.0 5.7 7.3 2.3 2.9 3.7 3.7 9.4 11.0 3.8 4.4 15.2 -8.4 11.3 -12.3 58.2 39.7 23.4 15.9

29/04/13 71 1.9 -61.4 -0.1 40.8 0.2 6.1 6.1 3.1 3.1 3.7 3.7 9.8 9.8 5.0 5.1 13.8 -10.9 - - 61.0 - 31.3 -

1/05/13 73 3.2 13.5 1.0 14.4 1.0 5.8 6.4 1.8 2.0 3.7 3.7 9.6 10.2 3.0 3.2 15.9 -5.5 - - 40.4 - 12.7 -

3/05/13 75 3.0 12.9 1.0 12.3 1.3 5.3 5.9 1.8 2.0 3.7 3.7 9.0 9.7 3.0 3.2 -1.8 14.1 - - 132.0 - 44.2 -

6/05/13 78 2.3 11.8 1.2 12.5 1.0 4.6 4.9 2.0 2.1 3.7 3.7 8.4 8.6 3.6 3.7 -7.5 7.6 -2.3 11.3 113.0 111.1 48.5 47.7

8/05/13 80 2.8 10.0 1.2 11.8 0.9 2.5 3.7 0.9 1.3 3.7 3.7 6.3 7.4 2.2 2.6 -5.2 5.5 -9.8 3.0 67.4 49.0 23.9 17.4

14/05/13 86 2.8 9.8 1.3 11.7 1.1 2.6 4.1 0.9 1.5 3.7 3.7 6.4 7.8 2.3 2.8 -2.1 9.0 -1.8 10.5 84.6 96.3 30.7 34.9

17/05/13 89 2.4 10.1 1.5 11.4 1.2 3.3 4.2 1.4 1.7 3.7 3.7 7.1 8.0 2.9 3.3 -1.7 10.7 -2.5 8.4 97.3 82.7 39.9 33.9

7/06/13 110 1.5 14.2 1.1 16.8 0.8 6.8 7.0 4.6 4.8 3.7 3.7 10.5 10.7 7.2 7.3 -41.2 2.5 -34.8 2.8 121.0 115.9 83.1 79.6

11/06/13 114 1.2 14.6 1.1 18.3 0.7 7.3 7.6 6.1 6.3 3.7 3.7 11.0 11.3 9.2 9.5 47.3 -4.0 -12.8 7.0 154.7 141.3 129.7 118.5

14/06/13 117 1.3 15.2 1.0 16.9 0.9 7.2 8.0 5.7 6.4 3.7 3.7 10.9 11.8 8.6 9.3 -19.5 7.3 -27.2 4.5 195.8 153.2 154.8 121.1

25/06/13 128 1.4 14.2 1.1 14.0 1.0 6.9 5.9 5.1 4.4 3.7 3.7 10.6 9.7 7.9 7.2 208.3 -0.9 29.8 -4.2 170.7 87.8 126.4 65.0

28/06/13 131 1.4 13.7 1.5 15.0 1.5 8.8 11.0 6.1 7.7 3.7 3.7 12.6 14.8 8.7 10.3 -11.7 8.0 -10.8 7.2 153.4 129.3 106.5 89.8

1/07/13 134 0.5 18.5 0.4 33.2 0.2 4.4 4.6 8.8 9.1 3.7 3.7 8.1 8.3 16.3 16.6 -9.3 7.8 -17.6 4.6 128.7 112.5 257.4 224.9

4/07/13 137 0.9 15.1 1.1 14.5 1.2 8.1 7.7 8.7 8.4 3.7 3.7 11.8 11.5 12.8 12.4 -16.4 5.3 -25.1 3.4 124.8 107.6 134.9 116.4

5/07/13 138 1.0 14.2 1.3 15.3 1.3 8.6 9.4 8.8 9.6 3.7 3.7 12.3 13.1 12.6 13.4 -13.2 6.6 -18.4 4.8 135.1 122.0 138.4 125.0

12/07/13 145 0.5 16.0 1.0 16.9 1.0 8.0 8.9 16.1 17.9 3.7 3.7 11.8 12.7 23.6 25.4 -42.2 2.7 -8.9 7.3 132.2 119.0 264.4 238.1

16/07/13 149 1.2 12.7 1.3 14.9 1.0 6.3 7.2 5.2 5.9 3.7 3.7 10.1 10.9 8.2 8.9 -6.8 15.3 -4.1 17.1 219.4 200.1 178.4 162.7

18/07/13 151 0.3 14.4 1.4 14.9 1.2 9.5 8.8 32.8 30.3 3.7 3.7 13.3 12.5 45.7 43.2 -3.5 14.1 -1.5 16.9 156.3 153.9 538.9 530.7

22/07/13 155 0.6 11.2 1.4 13.2 0.8 4.8 4.2 8.5 7.6 3.7 3.7 8.5 8.0 15.2 14.3 -10.7 6.6 -1.7 16.6 118.1 154.7 210.9 276.2

5/08/13 169 0.3 17.83 1.26 19.51 0.9 12.7 10.7 43.7 36.8 3.7 3.7 16.4 14.4 56.5 49.7 -16.6 6.6 -19.4 5.2 157.2 137.0 542.0 472.6

7/08/13 171 1.0 16.17 0.80 19.70 0.5 6.7 6.0 6.9 6.2 3.7 3.7 10.4 9.8 10.8 10.1 20.8 -5.5 -19.5 3.7 67.6 98.6 69.7 101.7

142.8 41.3 298.9 175.3

134.6 25.6 129.5 52.2

Calculation according the phase I of one-cycle DO profile

C(t) = A*(1-exp (-B*t))+Co*exp(-B*t) O2 consumption rate by OHOs = ((Cs-A)*B)+ralg-rnit

from phase Ia from phase Ib (mg O2/L.h) (mg O2/gVSS.h)from phase IIIa from phase IIIb

ralg (O2 prod rate by algae) = (A-Cs)*B

(mg O2/L.h) (mg O2/gVSS.h)

26.0 17.9

40.1 5.6 16.9 3.0

79.4 29.6 31.1 12.0

11.6 2.44 7.8 2.4 18.2 14.0

1.4

1

3

Calculation according the phase II of one-cycle DO profile

7.4 1.7 11.2 1.7

8.72

Date

Calculation according the phase III of one-cycle DO profile

4.1- 9.7 1.0 0.2-

4.9

Period DayVSS

(g/L) (mg O2/L.h) (mg O2/gVSS.h)

rnit(O2 consumption rate for nitrification) = Kla.Cs + ralg

- - 13.4 1.0 - 5.6 0.3

Kla.Cs

1.9 0.7 1.4

6.9 1.8 10.6 3.1

3.4 0.9

Appendix E

Light absorbance of mixed liquor at 550 nm

Date Chlo-a Absorbance

17-May 20.572 2.451

24-May 33.152 2.522

04-Jun 9.55 1.613

04-Jun 9.55 1.635

06-Jun 9.77 1.513

06-Jun 9.77 1.647

11-Jun 10.88 1.73

11-Jun 10.88 1.723

18-Jun 12.14 1.611

18-Jun 12.14 1.452

25-Jun 11.99 1.825

25-Jun 11.99 1.961

28-Jun 10.88 1.752

28-Jun 10.88 1.507

02-Jul 9.1 1.44

02-Jul 9.1 0.991

07-Aug 16.576 1.509

08-Aug 12.062 1.073

09-Aug 13.246 1.3