enhanced methane and hydrogen production in reverse...

91
i Thesis for the Degree of Doctor of Philosophy Enhanced Methane and Hydrogen production in Reverse Membrane Bioreactors via Syngas Fermentation Konstantinos Chandolias

Upload: others

Post on 08-Oct-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

i

Thesis for the Degree of Doctor of Philosophy

Enhanced Methane and Hydrogen production in Reverse Membrane

Bioreactors via Syngas Fermentation

Konstantinos Chandolias

Page 2: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

ii

ii

Copyright © Konstantinos Chandolias, 2019

Swedish Centre for Resource Recovery

University of Borås

SE-50190 Borås, Sweden

Digital version: http://urn.kb.se/resolve?urn=urn:nbn:se:hb:diva-21740

ISBN 978-91-88838-45-2 (printed)

ISBN 978-91-88838-46-9 (PDF)

ISSN 0280-381X, Skrifter från Högskolan i Borås, nr. 99

Cover page photo by Bruno Ramos Lara on Unsplash

Borås 2019

iii

Abstract

The increase in the waste production and the energy demand worldwide stimulates the

development of waste treatment processes, such as the anaerobic digestion. This biochemical

process converts organic substrates into biogas, with anaerobic microorganisms. However, some

types of substrates have low bio-degradability due to its recalcitrance or the presence of

inhibitors. This can be solved by the coupling of anaerobic digestion with gasification, a

thermochemical process that can convert organic substrates into syngas (H2, CO, and CO2)

regardless of the substrate´s degradability. Consequently, syngas can be converted into biogas

and other fermentative products via anaerobic digestion, in a process known as syngas

fermentation. In comparison to the catalytic conversion of syngas, syngas fermentation has

several advantages such as lower sensitivity to CO/H2/CO2 ratio and to syngas contaminants as

well as higher product specificity.

The main goal of this thesis was to improve the syngas conversion rate into CH4 and H2 by

addressing the cell washout, the cell inhibition by syngas contaminants, and the low gas-to-liquid

mass transfer, which are major challenges in syngas fermentation. For this purpose, a reverse

membrane bioreactor, containing a mixed culture encased in membranes, was used in various set

ups. The membranes were used in order to retain the cells inside the bioreactors, to protect the

cells against inhibitors, and to improve the gas holdup and gas-to-cell contact by decreasing the

rise velocity of syngas bubbles. As evident from the results, the cell washout was successfully

tackled during a continuous experiment that lasted 154 days. In addition, membrane bioreactors

fed with the syngas contaminants, toluene and naphthalene, achieved approximately 92% and

15% higher CH4 production rate, respectively, compared with the free cell bioreactors. In order to

improve the gas holdup and consequently the gas-to-liquid mass transfer of syngas, a floating

membrane bed bioreactor was set up. This bioreactor contained membrane sachets, filled with

inoculum that formed a packed floating membrane bed and achieved an increase of 38% and 28%

for the conversion rate of H2 and CO, respectively. Furthermore, the addition of a mixture of

heavy metals improved the production rates and yields during the syngas conversion into

fermentative H2.

Keywords: syngas fermentation; CH4; H2; cell washout; inhibitors; mass transfer

SVANENMÄRKET

Trycksak3041 0234

Page 3: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

iv

iv

PUBLICATIONS INCLUDED IN THE THESIS

I. Youngsukkasem, S., Chandolias, K., Taherzadeh, M. J. 2015. Rapid bio-methanation of

syngas in a reverse membrane bioreactor: Membrane encased microorganisms. Bioresource

Technology, 178, 334-340.

II. Westman, Y.S., Chandolias, K., Taherzadeh, M. J. 2016. Syngas Biomethanation in a Semi-

Continuous Reverse Membrane Bioreactor (RMBR). Fermentation, 2(8), 1-12.

III. Chandolias, K., Wainaina, S., Niklasson, C., Taherzadeh, M.J. 2018. Effects of heavy

metals and pH on the conversion of biomass to hydrogen via syngas fermentation.

Bioresources, 13(2), 4455-4469.

IV. Chandolias, K., Alan L., Izazi, N., Ylitervo, P., Wikandari, R., Milati, R., Taherzadeh, M., J.

2019. Protective effect of RMBR against syngas impurities. [submitted manuscript]

V. Chandolias, K., Pekgenc, E., Taherzadeh, M.J. 2019. Floating Membrane Bioreactors with

High Gas Hold-Up for Syngas-to-Biomethane Conversion. Energies, 12(6), 1046.

STATEMENT OF CONTRIBUTION

I. Responsible for the experimental work and partially responsible for the data analysis.

II. Responsible for the experimental work and partially responsible for the data analysis and

writing of the manuscript.

III. Conception of the idea and design of the experimental work. Partially responsible for the

experimental work, responsible for the data analysis and writing of the manuscript.

IV. Conception of the idea and design of the experimental work. Partially responsible for the

experimental work, responsible for the data analysis and writing of the manuscript together with

the co-authors.

V. Conception of the idea and design of the experimental work. Partially responsible for the

experimental work, responsible for the data analysis and writing of the manuscript.

v

PUBLICATIONS NOT INCLUDED IN THE

THESIS

VI. Chandolias, K., Pardaev, S., Taherzadeh, M.J. 2016. Biohydrogen and carboxylic acids

production from wheat straw hydrolysate. Bioresource Technology, 216, 1093-1097.

VII. Patinvoh, R.J., Osadolor, O.A., Chandolias, K., Sárvári Horváth, I., Taherzadeh, M.J. 2017.

Innovative pretreatment strategies for biogas production. Bioresource Technology, 224, 13-

24.

VIII. Chandolias, K., Richards, T., Taherzadeh, M.J. 2018. Chapter 5 - Combined

Gasification-Fermentation Process in Waste Biorefinery, in Waste Biorefinery, (Eds.) T.

Bhaskar, A. Pandey, S.V. Mohan, D.-J. Lee, S.K. Khanal, Elsevier, pp. 157-200.

Page 4: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

vi

vi

To Magda and Angelos, my sunshine

vii

PREFACE

This thesis is a part of the requirements for completing the Ph.D. program in Resource

Recovery at the University of Borås. The scientific results and articles that are presented were

produced under the supervision of Professor Mohammad Taherzadeh, Dr. Supansa Westman,

Professor Tobias Richards, and Professor Claes Niklasson.

Global issues, such as climate change, environmental deterioration, extinction of flora and

fauna, overconsumption, and dramatic increase in waste generation have been the main focus of

intensified scientific research projects in recent years. Syngas fermentation has several

advantages, such as rapid waste treatment, product variety, and improvement of the environment

as well as new job opportunities. In addition, the concept of coupling thermochemical and

biochemical processes can be a part of the sustainable solution to the increasing energy demand

and waste generation. This work focused on the syngas fermentation, which is the biochemical

conversion of syngas into fermentative biofuels and value-added chemicals. Several experiments

were conducted using different bioreactor designs. The background, the main results, and

conclusions are discussed in the following chapters.

Page 5: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

viii

viii

RESEARCH JOURNEY

Ever since I can remember, I have always been fascinated by chemistry and biology. My

father worked in the textile Industry, and I used to spend many summer days in the chemistry

room. While I was studying chemical engineering at the Aristotle University in Thessaloniki, the

issue of global pollution and climate change was becoming more popular. I thought that this will

be the major global issue in the future, we should make a difference, and change things for the

better; therefore, I started reading about renewable resources, while I did an internship in

environmental pollution and waste treatment. When I was accepted at the master’s program of

resource recovery-industrial biotechnology at the University in Borås, I was excited! After 2

years, I got the chance to start my research and teaching in the same area!

Good things come at a cost and my research journey was no exception. During my Ph.D.

studies, I tried to prioritize my tasks although as any Ph.D. understands, this is quite tricky.

Literature reviews, courses, experiments, conferences, teaching, administrative work, and

machines that broke down too often. At the same time, I was involved in several activities, such

as the working environment committee and the study visits from high school students, which

were very interesting experiences, although they required extra time and energy. However, it was

also an interesting period, as I tested my limits and interacted with many hard-working and smart

people. A main challenge at the beginning of my studies was to prove that I could use syngas, a

toxic and explosive gas, in a safe way; therefore, I put a lot of effort into writing risk declarations

and building a safe set up for the gas feeding inside a fume hood. Luckily, no accidents have

occurred during my experiments.

In the first year, I failed to finalize two experimental projects that lasted almost the whole

year. However, I had also gathered material from my previous experiments on the cell washout

from the bioreactors during repeated batch and continuous syngas fermentation. To achieve this, I

used a reverse membrane bioreactor, with my supervisor, Dr. Supansa Westman, who had come

up with this type of bioreactor. In this bioreactor, anaerobic cells are retained inside the

membrane sachets, which are immersed inside the liquid medium of the membrane bioreactor.

The main challenge in this work was to avoid membrane clogging because there was no way to

do a backwash in order to clean the membranes during the experiment. However, after trial and

ix

error with different set ups and because the produced biogas was pushing outwards through the

membrane pores, the membranes did not clog! The results were promising, and we managed to

publish the first two articles (paper I and paper II).

At the beginning of the second year, after discussions with my main supervisor, Professor

Mohammad Taherzadeh, we planned a new strategy for my research. Meanwhile, I started

writing a book chapter on the coupling of fermentation and gasification (paper VIII). This gave

me the chance to do a literature review in my area of studies, something that is difficult to do

when experiments are ongoing. During that year, the gas chromatograph, for which I was

responsible, stopped working and was fixed after 4 months. Then, I ran an experiment on the

conversion of straw hydrolysate into H2 and carboxylic acids. The results were not so promising;

therefore, the experiment was repeated in the third year, when it was finally published (paper

VI).

During the third year, I worked on the concept of using waste from gasification in order to

improve the fermentation yields and reduce the footprint of gasification. The results showed that

heavy metals found in gasification ashes could improve the H2 production, during syngas

fermentation (paper III). In addition, I was involved in the writing of a review paper on

innovative processes for biogas production (paper VII).

In the fourth year, I focused on the improvement of the low gas-to-liquid mass transfer,

which is a major bottleneck in syngas fermentation. I was trying to come up with an interesting

suggestion on how to use the reverse membrane bioreactor in order to overcome this limitation.

After several discussions with my supervisor and trials, I noticed that in a bioreactor that contains

enough membrane sachets and nourishment for the cells, a floating membrane bed is formed,

which could block the rising velocity of syngas and therefore increase the gas hold up. The idea

of passing syngas through several reverse membrane bioreactors, connected in line, failed.

However, syngas conversion rates were improved during a single pass through a reverse

membrane bioreactor (paper V).

In the fifth year, I tried to address the issue of the gas-cleaning requirement for raw syngas,

prior to fermentation. If gas cleaning is not required, then the economics of the overall process

can be improved. The idea was to use hydrophilic membranes in a reverse membrane bioreactor

Page 6: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

x

x

to block the contact of hydrophobic syngas contaminants with the cells. The main challenge was

which contaminants to choose and how they would react with the membranes. A long literature

review resulted in selecting some contaminants, which were tested. Finally, two of them, toluene

and naphthalene were picked for the experiment. The results proved the protective effect of the

membranes towards high concentrations of the studied syngas contaminants (paper IV).

During these 5 years, the failures were too many; however, they made the success-moments

even greater. The most valuable lessons were to learn to take criticism, turn failure into success,

and to collaborate with people from different cultures. I enjoyed being a Ph.D. student, and I feel

blessed for getting to experience this in Sweden. It also feels good to think that your work can

have an impact, and I hope that others will be helped and inspired by this study as I was from

other works.

xi

ACKNOWLEDGMENTS

I am grateful to the Swedish Research Council for the funding of this work.

I would like to express my gratitude to my main supervisor, Professor Mohammad

Taherzadeh for his guidance, availability, and for the opportunity to start with this Ph.D.

program. I also wish to thank Dr. Supansa Westman for her support and guidance during my first

year as a Ph.D. student and before that during my M.Sc. studies. I also thank Professor Tobias

Richards and Professor Claes Niklasson for their constructive discussions. Special thanks to

Assistant Professor Ilona Sárvári Horváth, for the encouragement, support, and warm welcome.

To the laboratory staff: Tomas, Kristina, Marlén, Haike, Faranak, and Jonas and to the

section staff especially to Peter, Patrick, Päivi, Jorge, Tomas, Solveig, Kamran, Magnus, Louise,

Jonas, Camilla, Tatiana, Jonas, Akram, and Dan; thank you for all your help and for creating a

nice working environment.

During my studies, I had the privilege to work with many bachelor’s, master’s, and Ph.D.

students from all around the world. I have learned a lot from you all, and I hope we will meet

again in the future! Thank you Karthik, Foluke, Wikan, Martin, Ram, Regina, Alex, Julius,

Mostafa, Maryam, Adib, Veronika, Farzad, Abas, Kehinde, Solmaz, Behnaz, Sunil, Amir, Luki,

Andreas, Rebecca, Sindor, Gürlu, Mohsen, Sofie, Madumita, Steven, Anette, Danh, Mukesh,

Pedro, Sajjad, Taner, Moein, Tuba, Hanieh, Babak, Eboh, Supri, David, and the rest of the Ph.D.

students that worked in the Swedish Centre for Resource Recovery. I also want to express my

gratitude to my master’s and bachelor’s students, Björn, Sara, Müge, Enise, Sara, and Karolin;

thank you for your hard work and efforts.

I feel very blessed to have my friends in Borås, and I feel that they have also helped me in

their own way in order to finish this thesis. Thank you Stamatis, Katerina, Argyro, Martha,

Iordanis, Vasiliki, Vicki, Thomas, Nelly, Spiros, Vaya, Stratos, Thanassis, Stavros, and Vasso.

My thoughts are always with my friends in Greece, especially to my best friend George;

thank you for the nice summer moments, I really needed them between my studies.

Page 7: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

xii

xii

To my lovely parents, Angelos and Rodi, my beloved brother Lampis, and my grandmother

Nikki. You have been a role model for me; thank you for your love, patience, and kindness. I

wish we could see each other more often.

To my wonderful wife Magda and my son Angelos. Without you, I could not make it.

Thank you for your love, and I hope that we will have more time together now. I love you!

xiii

Contents Abstract ......................................................................................................................................................... iii

PUBLICATIONS INCLUDED IN THE THESIS ...................................................................................................... iv

STATEMENT OF CONTRIBUTION ................................................................................................................... iv

PUBLICATIONS NOT INCLUDED IN THE THESIS ............................................................................................... v

PREFACE ....................................................................................................................................................... vii

RESEARCH JOURNEY .................................................................................................................................... viii

ACKNOWLEDGMENTS ................................................................................................................................... xi

NOMENCLATURE .......................................................................................................................................... xv

LIST OF FIGURES .......................................................................................................................................... xvi

LIST OF TABLES ........................................................................................................................................... xvii

Chapter 1 ....................................................................................................................................................... 1

Introduction ................................................................................................................................................... 1

1.1 Aim of this work .................................................................................................................................. 2

1.2 Thesis structure ................................................................................................................................... 3

1.3 Socioeconomic and ethical reflections ................................................................................................ 3

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

Anaerobic digestion ................................................................................................................................... 7

2.1 Biogas: an overview ............................................................................................................................. 7

2.2 Bio-H2: a renewable and sustainable product ................................................................................... 10

2.3 Basic principles of anaerobic digestion ................................................................................... .......... 11

2.3.1 Hydrolysis ................................................................................................................................... 12

2.3.2 Acidogenesis ............................................................................................................................... 12

2.3.3 Acetogenesis............................................................................................................................... 13

2.3.4 Methanogenesis ......................................................................................................................... 13

2.5 Waste with low degradability ............................................................................................................ 15

Chapter 3 ..................................................................................................................................................... 17

Thermochemical treatment for methane and hydrogen production ......................................................... 17

3.1 The gasification process .................................................................................................................... 18

3. 2 Types of gasifiers .............................................................................................................................. 21

3.3 Raw gas cleanup ................................................................................................................................ 22

Page 8: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

xiv

xiv

3.4 Catalytic conversion of syngas .......................................................................................................... 26

3.4.1 Hydrogen production ..................................................................................................... ............ 26

3.4.2 Methane production .................................................................................................................. 27

Chapter 4 ..................................................................................................................................................... 29

Coupling of thermochemical and biochemical process .............................................................................. 29

4.1 Metabolic activity for hydrogen and methane production ............................................................... 30

4.4 Operational parameters .................................................................................................................... 38

4.4.1 pH ............................................................................................................................................... 38

4.4.2 Temperature ............................................................................................................................... 38

4.4.3 Pressure ...................................................................................................................................... 39

4.4.4 Partial pressure........................................................................................................................... 39

4.5 Fermentation systems for hydrogen and methane production ........................................................ 40

4.5.1 Methane production .................................................................................................................. 40

4.5.2 Hydrogen production ..................................................................................................... ............ 43

4.6 The reverse membrane bioreactor ................................................................................................... 44

Chapter 5 ..................................................................................................................................................... 47

Improving syngas fermentation .................................................................................................................. 47

5.1 Cell washout ...................................................................................................................................... 47

5.2 Inhibitors ........................................................................................................................................... 50

5.2.1 Effect of heavy metals in anaerobic digestion ........................................................................... 50

5.2.2 Protective effect of rMBR against syngas contaminants ........................................................... 53

5.3 Gas-to-liquid mass transfer ............................................................................................... ................ 56

Chapter 6 ..................................................................................................................................................... 61

Conclusions and future work ....................................................................................................................... 61

6.1 Major experimental results and conclusions ................................................................................ .... 61

6.2 Future work ....................................................................................................................................... 62

6.2.1 Membrane material ................................................................................................................... 62

6.2.2 Bioreactor´s design ..................................................................................................................... 62

6.2.3 Effect of syngas contaminants and heavy metals ...................................................................... 62

References ................................................................................................................................................... 63

xv

NOMENCLATURE

ADP : Adenosine diphosphate

ATP : Adenosine triphosphate

CODH : Carbon monoxide dehydrogenase

ECH : Energy-conserving hydrogenase

ETP : Electron transport phosphorylation

FBEB : Flavin-based electron bifurcation

FCBR : Free cell bioreactor

Fd : Ferredoxin

ISR : Inoculum to syngas ratio

GHG : Greenhouse gas

SLP : Membrane bioreactor

NAD : Nicotinamide adenine dinucleotide

PBR : Packed bioreactor

PHA : Polyhydroxyalkanoates

PHB : Polyhydrobutyrate

PVDF : Polyvinylidene difluoride

RMBR : Reverse membrane bioreactor

SRB : Sulphate-reducing bacteria

SLP : Substrate level phosphorylation

Page 9: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

xvi

xvi

LIST OF FIGURES

Figure 2.1 Global biogas production in 2012 and its trend to 2022 9 Figure 2.2 Simplified scheme of the anaerobic stages 11 Figure 4.1 Schematic illustration of CO-oxidation, electron transfer, H2-formation and proton

transfer through the cell-membrane. CODH: CO dehydrogenase; Fd: ferredoxin; ECH: energy-conserving hydrogenase 33

Figure 4.2 The Wood-Ljungdahl pathway. H4F: tetrahydropholate; CoFeSP: corrinoid iron-sulphur protein 35

Figure 4.3 Metabolic pathway of aceticlastic methanogenesis. CoA: coenzyme A; H4SPT: tetrahydrosarcinapterin; HS-CoM: sulfhydryl coenzyme M; HS-CoB: sulfhydryl coenzyme B; CoM-S-S-CoB: heterodisulfide coenzymes CoM and CoB 36

Figure 4.4 Metabolic pathway of hydrogenotrophic methanogenesis. MFR: methanofuran; H4MPT: tetrahydromethanopterin; HS-CoM: sulfhydryl coenzyme M; HS-CoB: sulfhydryl coenzyme B; CoM-S-S-CoB: heterodisulfide coenzymes CoM and CoB 37

Figure 4.5 Schematic illustration of the reverse membrane bioreactor (RMBR) in the semi-continuous biomethanation process of syngas and organic substances. (a) Digesting sludge encased in PVDF membrane, (b) Organic and nutrient medium, (c) Syngas, (d) Peristaltic pump, (e) Gas outlet, (f) Warm water bath, (g) Effluent, (h) Flow meter, (i) Data analysis 45

Figure 5.1 Elimination of cell washout in a RMBR 48 Figure 5.2 Comparison of the CH4 production in RMBR and FCBR, during the semi-continuous

biomethanation of syngas and organic substances 49 Figure 5.3 Utilization of heavy metals from gasification ash during syngas fermentation 51 Figure 5.4 Relationships between H2 production activity, Ah (%) = (mol H2/mol H2 control) 100%

(a), H2 yield, YH = mol H2 prod./mol CO fed (b), initial medium pH and heavy metal concentrations (mg L-1) 52

Figure 5.5 Protective effect of membranes against syngas contaminants 54 Figure 5.6 The effect of naphthalene on the (a) CH4 production rate, mmol CH4 L-1 d-1; (b) CH4

production activity, (mol mol-1control) 100%; and (c) the theoretical naphthalene concentration, g L-1; in the FCBR and the RMBR 55

Figure 5.7 (a) Main resistances during the mass transfer of syngas from the gas phase to the site of reaction in the cells. Movement 1) in the gas bubble; 2) across the gas-liquid interfacial; 3) through the liquid film surrounding the gas bubble; 4) in the liquid bulk; 5) through the membrane pores; 6) inside the membrane; 7) across the liquid film surrounding the microbial cell; 8) through the cell membrane; 9) through the cell and end up in the site of reaction. (b) Movement of A through the interfacial boundary. CAL: concentration of A in the liquid phase; CAli concentration of A in the liquid boundary; CAGi concentration of A in the gas boundary and CAG concentration of A in the gas phase 57

Figure 5.8 Improvement of gas-to-liquid mass transfer rates in a floating MBR 59 Figure 5.9 Syngas biomethanation in a floating MBR, floating MBR/FCBR, PBR, and FCBR during

continuous syngas feeding. Consumption of (a) H2 and (b) CO and production of (c) CO2 and (d) CH4 in mmol·L-1·d-1) 60

xvii

LIST OF TABLES

Table 3.1 Principal reactions in gasification 19 Table 3.2 Composition of bottom and fly ashes from gasification of wood pellets, expressed as

mg kg-1 of the dry substance 20 Table 3.3 Composition of dry gas in oxygen and steam gasification 22 Table 3.4 Examples of feedstock contaminant levels 23 Table 3.5 Analysis of tar content in syngas derived from downdraft gasification with wood

pellets 24 Table 3.6 Analysis of syngas trace contaminants collected on filters after gasification of rice

hulls and wood chips. All concentrations are expressed in μg m-3syngas 25 Table 3.7 Gas cleaning requirements for different syngas applications 26 Table 4.1 Syngas fermenting bacteria 31 Table 4.2 Syngas fermenting methanogens 32

Page 10: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

1

Chapter 1

Introduction

Due to the increasing global waste generation, the climbing energy demand, and the

environmental deterioration from fossil fuels, there is an urgent need for renewable and

sustainable solutions. The anaerobic conversion of waste streams into biofuels, biochemicals, and

other bioproducts is a sustainable answer to the above challenges. During anaerobic digestion,

organic molecules, such as monosaccharides and amino acids, are broken down by

microorganisms, in the absence of oxygen. The main product of anaerobic digestion is biogas, a

mixture of CH4 and CO2, while other valuable products such as H2, acids, and bioplastics can also

be generated [1-3]. Although there are organic substrates that can be easily degradable by the

anaerobic cells, there are several types of potential substrate with low degradability, such as

lignocellulose. These substrates can be treated with chemical (e.g., acid) or biological (e.g.,

enzymatic) processes, which have a high cost.

Another more efficient method for the treatment of substrates with low degradability is

gasification. In this method, the feedstock is thermally degraded in the presence of an oxidizing

agent, usually air, oxygen, or steam. Gasification encompasses several advantages such as

feedstock flexibility and high conversion rates [4, 5]. The main product of gasification is syngas,

a gaseous mixture of mainly CO, H2, and CO2 with several applications, such as heat and power,

fuels (ethanol, H2, CH4), value-added chemicals (ammonia), and bioplastics [6]. During the past

few years, the anaerobic fermentation of syngas for the generation of biofuels and other value-

Page 11: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

2

2

added products, has gained broad interest in scientific, social, and industrial fields [7]. This

process has several advantages in comparison with the traditional catalytic processes, which

valorize inorganic/metal catalysts. More specifically, syngas fermentation is less sensitive to

syngas contaminants and to the H2/CO ratio; it does not require high temperature and pressure, it

is more product-specific, and no hazardous components are formed [8].

The coupling of gasification and fermentation is a relatively new technology, and several

challenges need to be tackled so that the efficacy of the overall process will be improved. Low

cell-density bioreactors are a common challenge in continuous anaerobic processes, where

productive bacteria are washed out of the bioreactors during the exchange of fresh and old liquid

medium. Another bottleneck is the high concentration of heavy metals inside the gasification-

derived ash, which poses an environmental and health hazard. Moreover, extended exposure to

high concentrations of syngas contaminants may be toxic for the anaerobic cells. Finally, a main

concern during syngas fermentation is the low gas-to-liquid mass transfer rates of syngas

components. The above challenges were the main focus of this thesis.

1.1 Aim of this work

The aim of this thesis was to overcome important challenges in syngas fermentation, such

as to eliminate the cell-wash out, to investigate the use of gasification-derived heavy metals in

fermentation, to study and address the cell sensitivity in syngas contaminants, and to improve the

low gas-to-liquid mass transfer of syngas components in bioreactors. For this purpose, a reverse

membrane bioreactor (RMBR) was employed and developed. The focus of the thesis can be

summarized in five steps:

1. To study syngas biomethanation in a RMBR, with high cell density, in batch conditions

(Paper I)

2. To operate a RMBR in continuous syngas biomethanation and report the effects of cell

retention (paper II)

3. To investigate the effect of different concentrations of heavy metals in syngas

fermentation for H2 production (paper III)

3

4. To report the beneficial and inhibitory concentrations of two syngas contaminants,

toluene and naphthalene, and the possible protective effect of the RMBR in batch and

continuous anaerobic digestion (paper IV)

5. To increase syngas holdup for higher gas-to-liquid mass transfer rates and thus higher

syngas conversion rates in a floating MBR with high gas holdup (paper V)

1.2 Thesis structure

This thesis is divided into 6 Chapters. In Chapter 1, the background, the aim, as well as the

socioeconomic and ethical reflections of this work are presented. The anaerobic digestion,

including the microbiology, digestion steps, and microorganisms are discussed in Chapter 2.

Moreover, the challenge of substrates with low bio-degradability is introduced. In Chapter 3,

gasification is presented as a sustainable treatment process for substrates with low bio-

degradability. Chapter 4 presents the syngas fermentation process, including biochemical

reactions, pathways, and bioreactors. The main challenges of syngas fermentation, which were

the focus of this thesis, including the cell washout, the utilization of heavy metals derived from

the gasification ashes, the effect of contaminants in the raw syngas, and the low gas-to-liquid

mass transfer rates of syngas components, are described in Chapter 5. Finally, the main

conclusions of this thesis and future work are discussed in Chapter 6.

1.3 Socioeconomic and ethical reflections

The ultimate goal in research is to create better living conditions for humanity. The way to

do that is by asking questions and answering with scientific data. During my research journey, I

have tried to correlate the importance and the effect of my studies to society´s well-being. At the

end of this work, I believe that my research can have a positive influence in society.

Although technology has rapidly evolved and improved our life standards, the dramatic

increase of waste and the rising energy demand create concerning environmental and social

challenges. Fossil fuels are the traditional resources used for energy; however, their use generates

serious pollution problems that are considered a main reason for our planet’s pollution and

climate change. The concept of circular economy with its three principles, to reduce, reuse, and

recycle materials, is becoming increasingly popular. Therefore, the use of abundant resources and

Page 12: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

4

4

the simultaneous reduction of waste are incredibly important for the present and future

generations.

This thesis is based on the waste-to-energy concept, with the combination of gasification

and fermentation for waste reduction and biofuel production. During the last few decades, there

has been an increasing interest toward biofuels, with global and regional policies. The production

of fermentative products such as biofuels and value-added chemicals can lead to a more

sustainable alternative to fossil fuels and petrochemicals. However, the use of first-generation

biofuels creates the ethical dilemma of food-to-energy conversion. The use of arable land, water,

and pesticides contradicts the concept of sustainability and raises global concerns. On the other

hand, the gasification-fermentation process uses renewable, abundant feedstock, such as forest

residues and municipal waste, for the production of second-generation biofuels. Moreover,

gasification and fermentation are two established industrial processes; thus, the infrastructure

exists. Moreover, intensification of the two processes will lead to more job opportunities in both

research and industrial projects.

Gasification offers rapid waste reduction with important social advantages. This creates

better living conditions, reduces the health effect and environmental pollution, and leaves more

space for human activities such as agriculture and farming. The reduction of waste affects

relatively large populations that live near landfills, especially in developing countries. These

populations depend on recycling or retrieving of valuable compounds from the waste, but they

are exposed to dangerous emissions and even explosions that take place in landfills because of

toxic and explosive gases formed inside the waste piles. The reduction of landfill waste also

decreases the threat of underground water pollution by leaching and the greenhouse gas (GHG)

effect by carbon dioxide and methane emissions. In addition, in many parts of the world, cooking

and waste reduction are still done by burning waste in the backyards, thus, creating

environmental and health threats. The development of a technology that converts various types of

waste, in a controlled and efficient way, into energy can dramatically minimize the uncontrolled

waste burning.

Syngas fermentation via anaerobic microorganisms can generate different products in the

concept of a biorefinery, thus, creating better economic opportunities. This work aimed to

achieve higher H2 and CH4 yields during syngas fermentation. These gases have numerous

5

applications, such as vehicle fuels, electricity generation, and production of chemicals. The

improvement of the yields could give a push on the biofuel market with higher revenues and thus

greater opportunities for new projects and jobs. In addition, except for syngas, other industrial

off-gases, with similar composition, could be used as a substrate. The reduction of CO and CO2

emissions could significantly reduce the industrial pollution in industrial areas. In addition, no

pathogens or other dangerous microorganisms are generated during fermentation. During this

work, only naturally occurring microorganisms, present in sewage sludge, were used as

inoculum. Therefore, there were no considerations about genetically modified cells and the

danger of them escaping into the environment.

This work was funded by public means; therefore, there has been a significant effort to

make the findings of the experiments publicly available by Open Access publications. In

addition, ethical norms, such as reliability, honesty, and respect for intellectual property were

followed in order to reassure the high publication standards.

Page 13: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

6

6

7

Chapter 2

Anaerobic digestion

Currently, the world is experiencing increasing rates of development, which aim for better

living conditions for societies. However, the drawbacks of this development are the

intensification in the use of fossil fuels and the increasing waste generation. This contributes

greatly to the global pollution and climate change. In order to address the above challenges,

renewable processes, such as anaerobic digestion, have gained considerable interest in the last

few decades. During anaerobic digestion, organic substrate is converted into biogas, by

microorganisms, under anoxic conditions. In addition, other intermediate products, such as

fermentative H2 can be obtained during anaerobic digestion.

2.1 Biogas: an overview

Biogas production technology is diverse, as it combines small and large scale plants and a

variety of possible substrates. In developing countries, biogas is commonly produced in small

domestic digesters for cooking or lighting, with food residues being the main substrate. In

developed countries, biogas production mainly takes place in larger scale industrial digesters for

electricity and heat production. In this case, wastewater sludge, food waste, manure, industrial

wastes, and agricultural residues are common microbial substrates. Biogas is usually produced in

digesters with operating temperature that ranges from 30 40 °C (mesophilic) to 50 60 °C

(thermophilic). Different anaerobic microbes dominate in different temperatures. In general,

Page 14: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

8

8

thermophilic conditions offer a faster degradation process and greater pathogen kill. The digestate

from thermophilic digesters can be used for land application with no restrictions, according to the

US Environmental Protection Agency (EPA) [9]. However, mesophilic conditions may be less

costly and easier to operate and maintain [9]. Wet substrate, containing less than 15% solids

content, is a more common substrate than dry feedstock, due to the fact that the wet feedstock can

be easily pumped in and out of the digester and be homogenized inside the digesters. Modern

digesters are equipped with other facilities that can deliver a broad range of products, from heat

generation to upgraded CH4 for vehicle fuel use. Other digesters are utilized for the treatment of

animal manure and human wastewater, and they are thus placed near farms and cities.

Biogas, consisting mainly of CH4 and CO2, has numerous applications, such as electricity

generation, combined heat and power plants (CHP), direct burning for energy generation and

cooking, injection into the natural gas pipeline, vehicle fuel, and fuel cells. In Europe, biogas is

largely used for generation of electricity, heat, or heat and power. The heat produced can be used

for the local facility demands or external users. In the case of biogas injection in the natural gas

network, biogas upgrade with removal of trace gases such as H2S, water, and CO2 is required

[10]. Biogas production from wastewater and landfill gas recovery, as well as biogas upgrading

into CH4 for vehicle fuel use or injection to the natural gas grid are gaining attention in several

countries [10].

Except from energy and fuel generation, biogas production results in environmental

benefits, such as decrease in water, soil, and air pollution [10]. Manure has traditionally been

used as a fertilizer in agriculture. However, this can cause environmental pollution due to

pathogen growth and CH4 and CO2 release in the atmosphere [11, 12]. In anaerobic digestion,

manure and other substrates are degraded in a controlled environment, resulting in a reduction of

odor and removal of pathogens that can pose health risks for humans and animals, while CH4 and

CO2 are used as the main products. In other words, the anaerobic digestion process results in the

reduction of Green House Gas (GHG) emissions. In addition, digestate from biogas production

can still be used as a biofertilizer containing similar nutrients to manure. This results in additional

economic benefits, while it reduces the use of chemical fertilizers, the nutrient runoff, and CH4

emissions [11-13].

9

Figure 2.1 Global biogas production in 2012 and its trend to 2022 [14]

The biogas market is expected to grow significantly in the following decade [15, 16] as

shown in Figure 2.1. The status of biogas market depends greatly on the price and availability of

fossil fuels. For example, the global oil crisis of 1973 led to an increased interest in biogas

production [17]. Although the oil prices decreased in 1985, and since 2015, there is still interest

in anaerobic digestion, due to environmental considerations [18]. Among the developed

countries, Germany is the leading country in biogas generation, which in 2010, generated 61% of

the total electricity produced from biogas in Europe [19]. The European commission legislates

incentives that aim for the production of biofuels and specifically biogas and biofertilizer. A new

European legislation includes a legally-binding EU-wide target of 32% for renewable energy by

2030 [20]. In 2015, biogas production in Europe represented half of the global biogas production

[10]. Furthermore, Europe is the world´s leading producer of bio-CH4 with 459 plants in 2015,

while most of the bio-CH4 production plants are in Germany (185 plants), UK (80 plants), and

Sweden (61 plants), in the same year [10]. In Asia, China and India are the leading countries in

biogas generation. China has the highest amount of domestic biogas plants, which in 2011

reached 41.68 million [21], while the large scale agricultural and industrial organic waste biogas

installations were 4,700 and 1,600, respectively [22].

9

Figure 2.1 Global biogas production in 2012 and its trend to 2022 [14]

The biogas market is expected to grow significantly in the following decade [15, 16] as

shown in Figure 2.1. The status of biogas market depends greatly on the price and availability of

fossil fuels. For example, the global oil crisis of 1973 led to an increased interest in biogas

production [17]. Although the oil prices decreased in 1985, and since 2015, there is still interest

in anaerobic digestion, due to environmental considerations [18]. Among the developed

countries, Germany is the leading country in biogas generation, which in 2010, generated 61% of

the total electricity produced from biogas in Europe [19]. The European commission legislates

incentives that aim for the production of biofuels and specifically biogas and biofertilizer. A new

European legislation includes a legally-binding EU-wide target of 32% for renewable energy by

2030 [20]. In 2015, biogas production in Europe represented half of the global biogas production

[10]. Furthermore, Europe is the world´s leading producer of bio-CH4 with 459 plants in 2015,

while most of the bio-CH4 production plants are in Germany (185 plants), UK (80 plants), and

Sweden (61 plants), in the same year [10]. In Asia, China and India are the leading countries in

biogas generation. China has the highest amount of domestic biogas plants, which in 2011

reached 41.68 million [21], while the large scale agricultural and industrial organic waste biogas

installations were 4,700 and 1,600, respectively [22].

Page 15: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

10

10

2.2 Bio-H2: a renewable and sustainable product

The current global H2 generation market is promising, and it was estimated to be larger than

100 billion USD in 2017 [23]. Between 2003 and 2008, there were reports of approximately 6%

yearly increase in the H2 sales [24]. Until 2026, the H2 market is expected to increase by more

than 2 times [23]. During the same period, Asia Pacific is expected to dominate the global H2

market [23]. The increase of the H2 market depends greatly on global trends for renewable

policies, on the status of technologies for producing/consuming H2, and the price of H2 and its

rivals, such as fossil fuels [25].

H2 is one of the most abundant components with a high value, several production processes,

applications, and properties. It has a high-energy content of 120 MJ/kg to 142 MJ/kg that is more

than 2.5 times higher than that of hydrocarbon fuels and releases only water during combustion

[26, 27]. The global H2 production is mainly used for ammonia manufacturing, with

approximately 50% contribution [24, 28]. Other typical applications of H2 are chemical

processes, lamps, vehicle fuel, laboratories, power-to-gas storage, and redox reactions (reductive

agent). The H2 use as a combustible fuel in stationary and transportation sector, such as in

internal combustion engines, rockets, fuel cell electric vehicles, and high-temperature industrial

furnaces is considered very promising [29, 30]. Moreover, H2 can be used as a blend with other

fuels, such as ethanol, in internal combustion engines [31].

H2 can be generated via several processes based on fossil fuels, such as reforming and

pyrolysis, or renewable sources. The main fraction of global H2 production is generated by CH4

reforming of natural gas, which utilizes fossil fuels [32]. Renewable processes for H2 production

include biomass treatment via biological or biochemical routes and water splitting. Water

electrolysis is a water splitting method with less environmental footprint, higher costs, and small

scale availability [27]. One biological process includes special bioreactors that use biocatalysts,

such as microalgae and phototrophic bacteria, for H2 production via photofermentation, which

have been studied [33]. In addition, fermentative H2 production by anaerobic digestion is

considered as a less expensive, less energy-demanding, and more eco-friendly process than other

production methods [34, 35]. Bio-H2 is an intermediate product in anaerobic digestion, in contrast

to biogas that is the end product of the process. The production of H2 and biogas is discussed in

more details in the following section.

11

2.3 Basic principles of anaerobic digestion

Figure 2.2 Simplified scheme of the anaerobic stages (adapted from [36])

Anaerobic digestion is a microbial process in which organic carbon is converted via

successive redox biochemical reactions into its most oxidized state (CO2), and to its most reduced

form (CH4) [18]. Anaerobic consortia contain a mixture of diverse microorganisms. In nature,

these cells thrive in different environments such as hot springs, wetlands, manure, and forest

sediments [37-40]. The anaerobic and facultative anaerobic microorganisms that take part in

anaerobic digestion have evolved a special symbiotic collaboration, called syntrophism due to the

small amount of energy available in CH4 generation. This means that different groups of bacteria

rely on each other for their metabolic activity, while together these bacterial groups show a

metabolic activity that no group could carry out on their own. The anaerobic digestion process

can be divided into four main stages: hydrolysis, acidogenesis, acetogenesis, and methanogenesis

as schematically presented in Figure 2.2.

11

2.3 Basic principles of anaerobic digestion

Figure 2.2 Simplified scheme of the anaerobic stages (adapted from [36])

Anaerobic digestion is a microbial process in which organic carbon is converted via

successive redox biochemical reactions into its most oxidized state (CO2), and to its most reduced

form (CH4) [18]. Anaerobic consortia contain a mixture of diverse microorganisms. In nature,

these cells thrive in different environments such as hot springs, wetlands, manure, and forest

sediments [37-40]. The anaerobic and facultative anaerobic microorganisms that take part in

anaerobic digestion have evolved a special symbiotic collaboration, called syntrophism due to the

small amount of energy available in CH4 generation. This means that different groups of bacteria

rely on each other for their metabolic activity, while together these bacterial groups show a

metabolic activity that no group could carry out on their own. The anaerobic digestion process

can be divided into four main stages: hydrolysis, acidogenesis, acetogenesis, and methanogenesis

as schematically presented in Figure 2.2.

Page 16: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

12

12

2.3.1 Hydrolysis

Hydrolysis is the first stage of anaerobic digestion where undissolved compounds, such as

carbohydrates, proteins, and fats are cracked down into water-soluble compounds. This

conversion is catalyzed by hydrolytic extracellular enzymes, which are excreted by hydrolytic

facultative and obligatorily anaerobic bacteria. Thermoanaerobium brockii is a representative

thermophilic, hydrolytic, bacterium. The presence of different hydrolytic bacteria depends on the

type of different substrates, and their task is to solubilize complex compounds, which will be

consequently transferred into the cells and further degraded by endoenzymes. The rate of this

stage depends on parameters, such as the production of enzymes, substrate structure, and

adsorption of enzymes on the substrate surface [41]. Hydrolysis can be the limiting stage of

anaerobic digestion when the solid and complex substrates are fed into the digester [42, 43].

Anaerobic bacteria, such as Clostridium and Bacteroides, pose an extra cellular multi-

enzyme complex, called cellulosome, for the decomposition of substrates such as carbohydrates

[44]. The typical structure of a cellulosome consists of large non-catalytic scaffoldin protein that

contains a carbohydrate-binding-module (CBM), surface layer homology (SLH), and several

cohesin domains [45]. Various enzymatic subunits are bound to the scaffoldin subunit via cohesin

and dockerin interactions. The SLH can bind on the bacterial cell wall, regardless of whether the

CBM binds on the substrate surface and the enzymatic subunits hydrolyze the substrate.

2.3.2 Acidogenesis

In acidogenesis, the products of hydrolysis are degraded by different facultative and

obligatorily anaerobic bacteria into compounds of one to five carbon units. Typical products of

acidogenesis are volatile fatty acids (VFAs), such as propionic, butyric, and valeric acid, and two

examples of acetogenic bacteria are the Butyribacterium methylotrophicum and the Clostridium

ragsdalei. Depending on the parameters, such as the type of substrate, the anaerobic conditions

and microbial consortia, others products, such as H2, CO2, alcohols, and ammonia can be

generated in this stage. Imbalances in this stage can affect the CH4 production as well. For

example, in case of high acid production (acidogenesis), the bioreactor may be overloaded and a

sudden drop in the pH may occur (souring). This can inhibit the methanogenic activity and

endanger the overall digestion process.

13

2.3.3 Acetogenesis

The main role of acetogenesis is the conversion of the products of acidogenesis, such as

volatile fatty acids and alcohols into H2, CO2, and acetate. Some examples of acetogenic bacteria

are Acetobacterium bakii, Acetoanaerobium noterae, and Acetitomaculum ruminis. The

production of excess H2 may result in the increase of H2 partial pressure. If this partial pressure

exceeds a limit, then the acetogenesis is not thermodynamically feasible. Therefore, a symbiotic

interaction between the H2 producing bacteria (acetogens) and the H2 consuming bacteria

(hydrogenotrophic methanogens) is vital for the digestion process [46]. In the presence of

sulphates, sulphate-reducing bacteria (SRB), such as bacteria of the genus Desulfobacter,

Desulfosarcina and Desulfovibrio, thrive in anaerobic digesters. These bacteria use H2 and

acetate and reduce sulphate into H2 sulphide. This leads to a concurrence of H2 use with the

hydrogenotrophic methanogens. In the presence of low acetate amounts and high sulphate

concentrations, the sulphate reducing bacteria obtain H2 and acetate more easily than the

methanogens.

2.3.4 Methanogenesis

In the last stage, CH4-forming microorganisms convert the products of acidogenesis and

acetogenesis, mostly acetate, H2, and CO2, into CH4. CH4 can also be formed by other organic

compounds, such as CO. Therefore, it is crucial that all substrate compounds have been

transformed to a form that can be treated by methanogens. This can be the slowest stage of

anaerobic digestion mainly because the methanogenic cells are more sensitive in comparison to

the microorganisms in the previous stages, they have lower growth rates, and they can consume

only a limited type of substrates. This means that for easily hydrolyzed substrates,

methanogenesis may be the limiting stage of the anaerobic digestion [47].

Methanogens can be classified according to their substrate utilization into three main

groups: the hydrogenotrophic, the acetotrophic, and the methylotrophic methanogens. The

hydrogenotrophic methanogens convert CO2 and H2 into CH4 and water. By this mechanism, they

maintain a low H2 pressure that is vital for the acetogenic and the acetotrophic bacteria. The

acetotrophic methanogens can convert acetate into CO2 and H2. The CO2 produced can be further

converted into CH4 by the hydrogenotrophic methanogens. Moreover, some acetotrophic

methanogens can convert CO and water into CH4 and CO2. The acetotrophic methanogens

Page 17: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

14

14

reproduce more slowly than the hydrogenotrophic methanogens and are negatively affected by

the high H2 partial pressure [47]. Therefore, the maintenance of low H2 pressure is vital for both

acetate and CH4 production. The methylotrophic methanogens convert water and substrates

containing the methyl group (CH3-), such as methanol and methylamine, into CH4, CO2, and

NH3.

2.4 Wastes-substrates for anaerobic digestion

The volumes of waste generated by human activity have been exponentially increasing after

the industrial revolution. According to the World Bank, the global waste generation will increase

by 70% on current levels by 2050. The waste generation per capita is affected by the national

gross domestic product (GDP); therefore, wealthier countries tend to produce more waste per

capita than poor countries [48]. However, as the GDP of giant countries such as India and China

is improving, the global waste generation is increasing.

Wastes can be categorized into solid, liquid, and gas streams including food, agricultural,

and forest residues, construction and electronic waste, municipal waste, and emissions from

transportation and industries. The composition of wastes can include glass, paper, metal, plastic,

lignocellulosic biomass, hair, and animal and human feces. Globally, efforts are made in order to

achieve a proper waste treatment, where the waste generation is reduced, reused, and recycled

(3Rs). The waste reduction can be achieved via a reduced consumption, less use of packaging

material and reuse of wastes, such as old textiles. Components, such as glass, paper, plastic, and

metal, can be recycled and reused. However, several waste streams are not recyclable and can be

treated for the generation of energy and other products, via other processes, such as anaerobic

digestion. The conversion of waste for the production of energy and value-added chemicals can

favor the concept of circular economy toward zero-waste societies, and lead to less dependency

on fossil fuels.

Various types of wastes including agricultural and forest residues, animal manure,

slaughterhouse waste, and municipal waste (e.g., household and sewage waste) can be used as a

substrate in anaerobic digestion. Historically, liquid waste streams, such as sewage sludge and

industrial wastewater have been fed in digesters. In the last few decades, the digestion of

municipal solid wastes and agricultural residues is gaining attention. The yield and composition

15

of fermentative products can vary greatly due to the content of carbohydrates, proteins, and lipids

in feedstocks. The elemental composition and the Buswell formula (Eq. 2.1) can be used in

practice in order to calculate the theoretical CH4 potential of each substrate. The theoretical CH4

potential of carbohydrates, proteins, and lipids is 0.42, 0.50, and 1.01 Nm3 CH4 kg-1 VS,

respectively [12]. Therefore, the theoretical methane potential of municipal solid waste, food

waste, and manure can be calculated at 180 350, 400 800, and 100 300 Nm3 CH4 ton-1 VS,

respectively [49], depending on their content in carbohydrates, lipids, and proteins.

CcHhOoNnSs + yH2O → xCH4 + nNH3 + sH2S + (c-x)CO2 (2.1)

Where: x = 1/8 (4c+h-2o-3n-2s)

For liquid substrates, with low solid content, the theoretical CH4 potential can be calculated

by the chemical oxygen demand (COD). In Eq. 2.2, one mole of CH4 needs two moles of oxygen

to oxidize CH4 to CO2 and water. Therefore, 0.35 L CH4 is produced per g COD substrate.

CH4 + 2O2 → CO2 + 2H2O (2.2)

A basic assumption for the above equations is that the substrate is completely degraded and that

the substrate utilization for biomass (microbial) growth is negligible [12, 50, 51]. The

calculations were made with the ideal gas formula, at standard conditions.

2.5 Waste with low degradability

Several waste streams, such as lignocellulosic material, textile residues, and keratin-rich

waste have relatively low degradability rates in anaerobic digestion. The protective structural

mechanism of these materials that prevents their degradation is called recalcitrance. One example

of recalcitrance is that the structure of some waste does not allow for efficient contact of the

substrate with enzymes and therefore, the digestion is inhibited. For instance, although

lignocellulose has high carbohydrate content, it cannot be easily degraded by anaerobic

microorganisms due to its lignin content [52]. Chicken feather and wool, with a high content of

protein, have low degradation rates because of their high content in keratin [53]. Moreover, there

are wastes that cannot be completely degraded or their degradation requires considerable amount

of time, for example, mixed landfill waste containing plastic, lignocellulose, food residues, etc.

Page 18: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

16

16

These materials contain degradable components; however, their diversity makes them

challenging substrates for anaerobic digestion. Another reason for the low degradability rate of

some wastes is their content of components that have antimicrobial activity. For example, the

fruit flavors, such as D-limonene in citrus waste can inhibit the anaerobic digestion process [54].

In order to convert wastes with low degradability into degradable substrates for anaerobic

digestion, various pretreatment methods can be employed. The available methods can be

categorized into mechanical, biological, chemical, and thermochemical methods. The mechanical

pretreatment includes substrate size reduction by grinding or milling and has a high-energy

demand. Enzymatic hydrolysis is a biological pretreatment that has low energy and cost-input

and long hydrolysis rates [53]. Alkalis, acids, oxidizing agents, and organic solvents are used in

chemical pretreatments. The efficiency of this type of pretreatment depends on the lignin content

and may lead to cellulose and hemicellulose losses, corrosion, cell inhibition, and loss of highly

volatile chemicals [53]. Another method to treat recalcitrant waste and waste containing

microbial inhibitors is via thermochemical treatment, such as combustion and gasification.

Especially in gasification, a large variety of wastes can be converted, at high temperatures and in

the presence of an oxidation agent, into a gas product that can consequently be converted into

fermentative products via anaerobic digestion. The process of gasification and the digestion of

the gaseous product are discussed in the following chapters.

17

Chapter 3

Thermochemical treatment for methane and

hydrogen production

Currently, thermochemical treatments, such as incineration (combustion), pyrolysis, and

gasification, of waste streams are applied in large scale. Incineration is commonly used for the

burning of feedstock with oxygen for heat and power generation [55]. Pyrolysis converts the

feedstock in the absence of oxygen into char and bio-oil [55]. Gasification is the conversion of

feedstock into a gas product. The thermochemical treatment of wastes offers advantages in

comparison to other types of waste treatment [5], such as the high conversion rates of solid waste

[4] and thus, the reduction of eventual uncontrolled CO2 and CH4 emissions from landfills [56].

Moreover, the thermochemical processes have wide feedstock flexibility, including waste with

low degradability rate, such as forest residues. Another advantage is the destruction of pathogenic

microbes because of high operational temperatures [57]. The main advantage of gasification in

comparison to other thermochemical processes is the generation of the gaseous product that has a

high value and numerous applications. In addition, several operating conditions of the

gasification can be altered upon demand, creating a flexible process [58, 59]. This chapter

focuses on the process of gasification, equipment, product applications, and challenges, such as

the toxicity of residuals, the contaminants in raw syngas, and the catalytic conversion of syngas

into CH4 and H2.

Page 19: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

18

18

3.1 The gasification process

During gasification, low-value feedstock is converted into high value products at high

temperatures and the presence of oxidizing media. At temperatures below 1000 °C, a gas blend of

mainly CO, H2, CH4, CxHy aliphatic hydrocarbons, tars, CO2, and water is produced [60]. At

temperatures above 1200 °C, syngas is produced, which consists mainly of CO, H2, CO2, together

with a small amount of CH4, nitrogen, and contaminants, depending on the operating conditions

[60].

Gasification consists of four main steps: drying of feedstock, pyrolysis, oxidation, and

reduction. The order of these steps depends on the gasifier set-up. Drying takes place in order to

remove moisture, at temperatures above 100 °C. Pyrolysis is the thermal decomposition of

feedstock in the absence of an oxidation agent, and occurs at 150 700 °C. The feedstock

moisture ranges between 5–35% and generates steam when evaporated. Pyrolysis gases, such as

CO2, CO, H2, CH4, and water vapor together with char, tar, and volatile compounds are released

in this stage. The use of oxygen as an oxidation agent, although it is more expensive than air, is

often preferred in order to avoid high amounts of nitrogen [61]. Table 3.1 shows important

reactions that take place during gasification. In the oxidation zone, the oxygen can react with

solid carbonized fuel, producing CO or CO2 as shown in Eq. 3.1 and Eq. 3.2. In Eq. 3.3, the

water vapor introduced with air or produced by drying or pyrolysis reacts with the hot carbon,

according to the reversible heterogeneous water gas reaction. In the reduction zone, several

reactions take place in the absence of oxygen. Principal reduction reactions are the water, gas and

Boudouard reaction (Eq. 3.3 3.4). In addition, other important reduction reactions are the water

shift (Eq. 3.5) and the methanation reaction (Eq. 3.6). The water shift reaction is exothermic

when water is in surplus. When the reactions in Eq. 3.7 and Eq. 3.8 occur, heat is produced;

however, the heat value is reduced.

19

Table 3.1 Principal reactions in gasification [62]

Reaction ΔH, kJ mol-1

(25 °C, 1 atm)

Equation

C + 0.5O2 → CO -123.1 (3.1)

C + O2 → CO2 -393.8 (3.2)

C + H2O ↔ CO + H2 +118.5 (3.3)

CO2 + C 2CO +159.9 (3.4)

CO + H2O CO2 + H2 -40.9 (3.5)

C + 2H2 CH4 -87.5 (3.6)

CO + 0.5O2 → CO2 -283.9 (3.7)

H2 + 0.5O2 → H2O -285.9 (3.8)

After gasification, a relatively small amount of feedstock that has not been converted into the gas

phase is left in the gasifier in the form of ash. The elemental composition of ash that can contain

carbon, minerals, and metals can include Ca, K, P, Cu, Zn, Mn, Fe, and Mg [63] (Figure 3.2).

The residual phase is classified according to its carbon content, to char (high content) and ash,

which contains high concentrations of metals and minerals [63]. The char is commonly used in

combustion for heat generation, while the ash can be used as a construction material after

pretreatment [63]. For example, a study that investigated alternative uses of gasification ash

stated that the ash could be used as a component in bricks without pretreatment [64]. However,

there is a lack of research on the efficient uses of gasification-derived ashes in the literature. A

main challenge with incineration ashes used as construction material in roads is leaching, which

depends greatly on the weather conditions [65]. Another potential use of these ashes is the

anaerobic digestion. Components of the ashes, such as the heavy metals, can improve the yields

of fermentative products [66]. This was the focus of a part of this thesis (paper III), and it is

discussed in more details in chapter 5.

Page 20: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

20

20

Table 3.2 Composition of bottom and fly ashes from the gasification of wood pellets, expressed as mg kg-

1 of the dry substance (adapted from [67])

Bottom ash Fly ash

Element 550 °C 800 °C 550 °C 800 °C

Ca 308.0 365.3 309.2 354.7

K 108.0 47.1 48.3 33.8

Mg 32.6 37.8 34.3 35.4

Fe 17.8 23.1 33.3 51.4

Al 11.7 19.6 16.0 17.5

Na 7159.0 8929.0 6212.0 5932.0

Mn 3625.0 4362.0 4979.0 5055.0

Phosphate 36.7 44.7 41.0 44.3

Sulphate 17.1 19.9 18.5 21.6

Si 73.7 n.d. 82.9 n.d.

Zn 466.0 283.0 1416.0 1700.0

B 397.0 464.0 546.0 538.0

Cu 159.0 353.0 150.0 178.0

Cr 111.0 234.0 252.0 455.0

Ni 92.0 104.0 143.0 165.0

Pb 7.0 8.0 47.0 53.0

Cd 0.8 1.1 6.3 6.3

Co 9.0 14.1 n.d. n.d.

21

3. 2 Types of gasifiers

There is a wide variety of commercial gasifiers for waste gasification, with the majority of

them employed for heat and power generation. The main difference in these gasifiers is the

altered parameters, such as design and the operating conditions. The feedstock entry can be done

from the lower side, bottom, or upper side, and the oxidizing agent is usually oxygen, air or

steam. Typical gasifier designs are downdraft and updraft gasifier, bubble, circulating and dual

fluidized bed gasifier, and plasma gasifier [68]. The downdraft gasifier is commonly used for

small or medium applications, with the feedstock added at the top of the gasifier and landing on a

grate. In the updraft gasifier, the feedstock is added at the top of the gasifier, more diverse

feedstock can be treated, and the gasifier has high-energy efficiency. A main difference in these

gasifiers is that the oxidizing media are fed at the bottom of the updraft gasifier. In entrained flow

gasifiers, the feedstock and the oxidation agent are moving in the same direction. The main

feature of the fluidized bed gasifiers is that the feedstock and bed material are levitated by the air

stream that is supplied at the bottom of the gasifier. Finally, in plasma gasification, the feedstock

is gasified by electrically generated plasma at high temperatures (1500 5000 °C) and

atmospheric pressure, in the presence of oxidizing media for the production of high quality

syngas. Autothermal gasifiers provide the heat required for the gasification reactions, in means of

partial oxidation inside the gasifier. This is the main advantage of this type of gasifiers in

comparison to the allothermal gasifiers, where the heat production and heat consumption are

done in separate facilities [69].

Another categorization of gasifiers is by the type of oxidizing media, such as the oxygen

blown and the steam blown gasifiers. The composition of the gas compounds differs greatly when

oxygen or steam is used as an oxidation agent during gasification (Table 3.3). During oxygen

gasification, the CO2 and water produced during the combustion, take part in the chemical

reactions. On the other hand, higher amounts of H2, which derives from the steam, are obtained in

the steam gasification. The tar content is highly dependent on the operating temperature. For

example, in entrained gasification with operating temperature above 1000 °C, a low amount of tar

is produced.

Page 21: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

22

22

Table 3.3 Composition of dry gas in oxygen and steam gasification (adapted from [69])

Compound (content) Oxygen gasification

(entrained flow)

Oxygen gasification

(fluidized bed)

Steam gasification

CO (vol %) 40–60 20–30 20–25

CO2 (vol %) 10–15 25–40 20–25

H2 (vol %) 15–20 20–30 30–45

CH4 (vol %) 0–1 5–10 6–12

N2 (vol %) 0–1 0–1 0–1

LHV (MJ Nm-3) 10–12 10–12 10–14

Tar content (g Nm-3) <0.1 1–20 1–10

3.3 Raw gas cleanup

Typical gas contaminants consist of char, tars, nitrogen (NH3, HCN, etc.), sulfur (H2S,

COS, etc.), H2 (HCl, HF, etc.), trace metals (Na, K, etc.) S, alkali, Cl, and particulate matter [70].

The composition of these contaminants varies greatly and is significantly influenced by the

feedstock contaminants and the gasification operating conditions [70]. Examples of feedstock

contaminant composition, tar content in syngas, and trace syngas contaminants are shown in

Table 3.4, Table 3.5, and Table 3.6, respectively.

23

Table 3.4 Examples of feedstock contaminant levels [71-73], (adapted from [70])

Contaminant Wood Wheat straw Coal

Percent by mass

Sulfur 0.01 0.2 0.1–5.0

Nitrogen 0.25 0.7 1.5

Chlorine 0.03 0.5 0.12

Ash (Major Components) 1.33 7.8 9.5

K2O 0.04 2.2 1.5

SiO2 0.08 3.4 2.3

Cl 0.001 0.5 0.1

P2O5 0.02 0.2 0.1

The cleanup of raw syngas is essential prior to the syngas application in downstream

processes [74]. Table 3.7 shows examples of gas cleaning requirements for different syngas

applications. The gas cleanup processes can be classified according to the temperature range into:

hot gas cleanup (HGC), warm gas cleanup (WGC), and cold gas cleanup (CGC) [70]. Cold gas

cleanup takes place at ambient temperatures, where water spray is used and water condenses in

the outlet. Contaminants are either washed away with the condensed water or act as condensation

sites for water. Hot gas cleanup occurs at temperatures as low as 400 °C, while only few hot

cleanup processes manage to operate at temperatures higher than 600 °C. During this type of gas

cleanup, syngas is purified by contaminants, such as alkali compounds [75]. Warm gas cleanup

takes place at temperature ranges between the water boiling point and 300 °C, which allows

contaminants, such as ammonium chloride, to condensate. In addition, at high temperature gas

cleanup processes, candle filters for removing solid contaminants and sorbents for removing fluid

contaminants are employed [76]. An alternative way to treat syngas containing contaminants, to

avoid syngas cleaning requirements, was investigated in this thesis. The protective effect of

membrane encasement of cells against syngas contaminants (tars) was studied in a RMBR

Page 22: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

24

24

(Paper IV). As shown in Figure 3.4, a long list of tars can be found in syngas in various

concentrations; however, in this work, due to time and resource restrictions, two common syngas

contaminants, toluene and naphthalene, were investigated [77]. The results of the experimental

work are discussed in more details in chapter 5.

Table 3.5 Analysis of tar content in syngas derived from downdraft gasification with wood pellets (adapted from [77])

Compound Concentration,

mg Nm-3

Compound Concentration,

mg Nm-3

Toluene 76.8 198.3 2-Vinylnaphthalene 0.4 6.7

o/p-Xylene 9.3 111.6 Furfural 0.0 4.0

Naphthalene 62.3 126.1 Naphthalene, 1,8-dimethyl- 0.6 3.6

Phenol 6.9 67.2 Naphthalene, 1,5-dimethyl- 0.0 3.6

Styrene 21.0 65.1 Dibenzofuran 0.4 3.4

Indene 15.7 55.8 alpha-Methylstyrene 1.5 3.1

Ethylbenzene 2.5 25.0 2-Ethyltoluene 0.6 3.0

Phenol, 3-methyl- 1.3 25.4 Benzene, 1,2,3-trimethyl- 1.4 2.4

Benzofuran 8.5 24.9 Phenol, 2,4-dimethyl- 0.0 2.4

Biphenylene 7.1 22.2 Acenaphthene 0.3 2.1

Benzofuran, 2-methyl- 0.0 23.8 Phenol, 3,5-dimethyl- 0.0 1.9

m-Methylstyrene 6.6 18.8 Naphthalene, 2,3-dimethyl- 0.0 1.4

Naphthalene, 2-methyl- 5.1 16.2 Phenol, 3-ethyl- 0.0 1.3

Naphthalene, 1-methyl- 5.9 14.6 Phenol, 4-ethyl- 0.0 1.0

Biphenyl 2.6 10.1 Naphthalene, 1,8-dimethyl- 0.0 0.8

Phenol, 2-methyl- 0.5 8.9 Total tar 340 680

25

Table 3.6 Analysis of syngas trace contaminants collected on filters after gasification of rice hulls and wood chips (adapted from [78]). All concentrations are expressed in μg m-3syngas.

Feedstock Feedstock

Compound Rice hulls Wood chips Compound Rice hulls Wood chips

PM2.5 12.74 11.88 Na 1.45 1.9 Organic carbon

18.43 21.69 Mg 0.09 0.33

Elemental carbon

0.12 1.57 Al 0.87 0.18

Cl- <0.01 <0.01 Si 0.12 0.02

NO3- 0.87 0.55 S <0.01 <0.01

SO4-2 0.40 0.35 Cl 0.37 0.07

NH4+ 0.93 0.90 K 0.44 0.05

Na+ 0.54 <0.01 Ca 0.04 0.06

K+ 0.67 0.12 Fe 0.01 0.08

NH3 35.35 24.17 Zn 0.01 0.07

HCl n.d. n.d. Br <0.01 0.03

HNO3 2.62 3.08 Pb <0.01 0.01

SO2 5.74 <0.01 Others (sum) 1.22 1.80

H2S 0.53 <0.01

Page 23: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

26

26

Table 3.7 Gas cleaning requirements for different syngas applications [79-82] (adapted from [70])

Contaminant Application

Internal combustion engine

Gas turbine Methanol synthesis

Fischer-Tropsch synthesis

Particulate <50 mg m−3 <30 mg m−3 <0.02 mg m−3 n.d.*

Tars (condensable) <100 mg m−3 <0.1 mg m−3

Sulfur <20 μL L−1 <1 mg m−3 <0.01 μL L−1

(H2S, COS)

Nitrogen <50 μL L−1 <0.1 mg m−3 <0.02 μL L−1

(NH3, HCN)

Alkali <0.024 μL L−1 <0.01 μL L−1

Halides (primarily HCl)

1 μL L−1 <0.1 mg m−3 <0.01 μL L−1

* n.d. = not detectable

3.4 Catalytic conversion of syngas

The main applications of syngas include the production of ammonia (50%), H2 (25%),

methanol, and Fischer-Tropsch products [83]. In the last decade, R&D processes have mainly

focused on the production of transportation fuels [69]. The production processes of H2 and CH4

from syngas via catalytic conversion are further discussed in the following paragraphs.

3.4.1 Hydrogen production

Syngas can be converted into H2 via the water gas shift (Eq. 3.4) and steam reforming (Eq.

3.5) reactions [61]. Steam reforming is a chemical synthesis process for the production of H2 and

CO from hydrocarbons, at a high temperature and pressure and the presence of a catalyst. Most of

the global H2 production is generated via the steam reforming of natural gas [84]. The CO can be

converted by catalysts, such as iron and chromium, at high temperatures (400 500 °C) [85].

Relatively pure H2 can also be produced by the steam reforming process followed by water gas

27

shift reaction [85]. The CO content can decrease to approximately 2% and 0.2% after the water

gas shift reaction at high and low temperature, respectively [85].

Another way to produce H2 from natural gas is the partial oxidation, where CH4 and other

hydrocarbons react with a limited amount of oxygen, to produce H2, CO, and heat (Eq. 3.9). In

comparison to steam reforming, this process is faster, requires a smaller reactor and gives off

heat. On the other side, more H2 per unit of fuel is produced in steam reforming. The CO

produced can consequently be removed by the water gas shift reaction (Eq. 3.5).

CH4 + 0.5O2 ↔ CO + 2H2 + (heat) (3.9)

During the water gas shift reaction, metals and metal oxides are utilized as catalysts.

Copper-based catalysts are preferred at lower temperatures, while iron oxide-based catalysts are

used at high temperatures [86]. These catalysts are sensitive to impurities. For example, the high

temperature catalyst Fe/Cr (350–500 °C) can be deactivated at sulphur concentrations higher than

200–250 ppm, and the low temperature catalyst Cu/Zn (150–300 °C) can be deactivated at

sulphur concentrations as low as 1–10 ppm [87, 88]. Other catalysts based on cobalt and

molybdenum sulphides are tolerant to sulphur [85]. The CO2 after the water gas shift reaction can

be removed by water absorption [89].

3.4.2 Methane production

CH4 is the main component of synthetic natural gas (SNG), a gas generated by coal or

biomass that can substitute natural gas [90]. The methanation takes place after H2 and CO react in

the presence of catalysts, such as Ni, according to the reaction below:

CO + 3H2 → CH4 + H2O (3.10)

The reaction can occur even at atmospheric pressure, although higher pressure is preferred.

A H2/CO ratio of approximately 3.0 is required, which can be achieved by the water gas shift

reaction. In some types of reactors, such as fluidized beds, methanation can occur in parallel with

the water gas shift reaction; thus, there is no need for additional adjustment of the H2/CO ratio

[61]. An important limitation of this process is that the Ni catalysts are sensitive to sulphur

poisoning. In syngas methanation, removal of CH4 from syngas is not required. This is an

Page 24: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

28

28

important advantage for the production of bio-SNG. Both fixed and fluidized bed reactors have

been employed for syngas methanation [90]. A main concern of this process is the heat removal,

which can be addressed by a series of fixed bed reactors with gas cooling [90]. On the other hand,

the use of fluidized bed offers better heat transfer and removal capabilities [91]. However, low

rates and product selectivity in fluidized beds, due to gas and solids backmixing, are important

challenges [91].

An alternative route to the catalytic conversion of syngas to H2 and CH4 is the syngas

fermentation, where microorganisms are utilized as biocatalysts. This combination of

thermochemical and biochemical processes is considered to become more cost effective and more

robust. The technology of syngas fermentation and its advantages in comparison to the catalytic

conversion processes are discussed in the next chapter.

29

Chapter 4

Coupling of thermochemical and biochemical process

The biochemical process of anaerobic digestion can convert organic feedstock into H2 and

CH4. However, the anaerobic digestion cannot efficiently treat a large fraction of potential

feedstock with low degradability, such as lignocellulosic material. This can be overcome by

combining thermochemical and biochemical treatments. Gasification can treat feedstocks

regardless of its low degradability, and the syngas produced can be further anaerobically digested

via syngas fermentation. The coupling of thermochemical and biochemical processes for the

production of fermentative products is a promising strategy with several advantages [92, 93].

Gasification offers high conversion efficiency and a broad range of feedstock, regardless of their

biodegradability, while the biomethanation of syngas is a flexible multipurpose method that can

be used for converting biomass-derived syngas and industrial off-gases, rich in CO and CO2 [94].

In addition, an in-situ biogas upgrading can take place by coupling of an external H2 supply to the

syngas biomethanation bioreactor [95]. In comparison to the combined thermochemical and

catalytic conversion, syngas fermentation has a higher biocatalyst specificity, lower energy costs,

lower sensitivity of biocatalysts to impurities and to the H2/CO ratio, while ambient pressures and

temperatures are required [96, 97]. Moreover, the biological catalysts in the syngas fermentation

can generate a larger variety of products, and the biochemical reactions are irreversible.

Syngas fermentation was the main focus of this thesis. During this process, H2 and CO can

be converted into fermentative products by anaerobic microorganisms that use syngas as a carbon

Page 25: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

30

30

and energy source. Some of the main products that have been investigated in recent studies are

ethanol, acetone, butanol, 2.3-butanediol, and polymers, with ethanol being the most popular [8,

96, 98-100]. The effect of parameters, such as pH, temperature, syngas contaminants, and gas

partial pressure has been investigated [101-105]. In addition, several works studied the effect of

bioreactor design in syngas fermentation [94, 106]. The slow gas-to-liquid mass transfer of

syngas components is a main challenge; thus, it has been investigated in different operating

conditions, bioreactor designs, and cultures. For example, a study reported that the CO mass

transfer improved by using an external hollow fiber membrane for gas diffusion [107], while

another study achieved improved mass transfer of syngas in a monolithic biofilm reactor [108]. In

addition, the effect of some syngas contaminants on the microbial activity has been studied [104]

in correlation with the use of pure and mixed cultures. Research has shown that the use of a

mixed culture could be more beneficial in large-scale syngas fermentation due to no requirements

for sterile conditions and due to the higher tolerance and adaptation to syngas contaminants [8,

109, 110].

The technology of syngas fermentation for H2 and CH4 production is still not commercial

because of the high sale prices that are required to support it [110, 111]. The scaling-up of the

process can be achieved by actions such as the coupling of syngas fermentation with already

active gasification activities and achieving higher productivity yields in syngas fermentation

[110]. One of the most important steps toward the commercialization of syngas fermentation is

the local and global policies that aim to replace fossil fuel resources though sustainable processes.

4.1 Metabolic activity for hydrogen and methane production

Syngas can be converted into biofuels and chemicals by several anaerobic microorganisms

with diverse sizes, shapes, and inhabitants. Syngas fermenting microorganisms also have diverse

catabolic metabolisms and can generate different products, such as carboxylic acids, alcohols, H2,

CH4, and CO2. Several of the known strains that have been isolated are gram-positive, and it is

mainly carboxydotrophic hydrogenogens that grow chemolithotrophically during the conversion

of CO and water into H2 and CO2, at 55–80°C [26]. Table 4.1 and Table 4.2 show known

acetogens and methanogens that can ferment syngas components. The vast majority of the known

syngas-fermenting microorganisms are acetogens.

31

Table 4.1 Syngas fermenting bacteria (adapted from [68])

Microorganisms pH T,

°C Substrate Products Reference

Acetobacterium

woodii (wild) 7 30

CO2-H2,

PH2 = 1700 mbar Acetate [112]

Butyribacterium

Methylotrophicum

(adapted on CO)

7.3 37 CO:CO2 =

70:30

Acetic &

butyric

acid

[113]

Clostridium

aceticum (wild) 7.1–8.9 30

CO:H2:Ar =

78:4:18 Acetate [114]

Clostridium

ljungdahlii (wild) 4–6 37 CO:CO2:H2:Ar=

55:10:20:15

Acetate,

ethanol [115]

Clostridium

ragsdalei, P11

(wild)

5–6 37 CO:CO2:H2:N2=

20:15:5:60

Ethanol,

acetic

acid, 1-

butanol

[116]

Carboxydothermus

hydrogenoformans

(wild)

6.8–7 70 CO, 100% Acetate,

methanol,

ethanol

[117]

Citrobacter sp Y19

(wild) 7 30

CO: air+organics

=

2.5:97.5 (v/v)

H2 [118]

Clostridium drakei

(wild) 6–7 25

30 H2/CO2, CO/CO2 Acetate [119]

Page 26: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

32

32

Table 4.2 Syngas fermenting methanogens (adapted from [68])

Microorganisms pH T,

°C Substrate Products Reference

Methanosarcina

barkeri (adapted on

CO)

7.4 37 CO, 100% CH4 [120]

Methanosarcina

acetivorans strain

C2A (adapted on CO)

7 37 CO, 100% Acetate,

formate,

CH4

[121]

Methanothermobacter 7.4 65 CO: H2/CO2= CH4 [122]

Thermococcus

onnurineus NA1 6.1–6.2 80 CO, 100% H2 [123]

The production of H2 and CH4 can take place via different biochemical reactions [124],

depending on the syngas composition, existence of co-substrates, operating conditions, and type

of consortia. The conversion of CO can be done by carboxydotrophic hydrogenogens via the

biological water gas shift reaction (Eq. 3.1), by carboxydotrophic acetogens (Eq. 3.2) via the

Wood Ljundgahl metabolic pathway, and by carboxydotrophic methanogens (Eq. 3.3). However,

the CH4 production via carboxydotrophic methanogenesis is considered rare in syngas

biomethanation [110]. The conversion of H2/CO2 is done by homoacetogens (Eq. 3.4) via the

Wood Ljundgahl metabolic pathway by hydrogenotrophic (Eq. 3.5) and aceticlastic (Eq. 3.6)

methanogens. During syngas fermentation, ATP can be produced through different ways, such as

substrate level phosphorylation (SLP), electron transport phosphorylation (ETP), and electron

bifurcation [125-127]. The biochemical reactions, the pathways, and energy conservation are

further discussed in the following paragraphs.

The biochemical reactions are presented with the standard change of Gibbs free energy

[110, 124].

CO + H2O → CO2 + H2, ΔG0 = -20 kj/mol (3.1)

33

4CO + 2H2O → CH3COOH + 2CO2, ΔG0 = -165.4 kj/mol (3.2)

4CO + 2H2O → 3CO2 + CH4, ΔG0 = -210.9 kj/mol (3.3)

2CO2 + 4H2 → CH3COOH + 2H2O, ΔG0 = -104.6 kj/mol (3.4)

CO2 + 4H2 → CH4 + 2H2O, ΔG0 = -135.6 kj/mol (3.5)

CH3COOH → CH4 + CO2, ΔG0 = -31 kj/mol (3.6)

The carboxydotrophic hydrogenogens conserve energy via the H2 formation (Figure 4.1)

during the biological water gas shift reaction. The main enzyme that catalyzes this reaction is

CO-dehydrogenase (CODH), which oxidizes CO into CO2. An iron-sulfur protein called

ferredoxin (Fd) transfers the released electrons to an energy-conserving hydrogenase (ECH),

where they are bound with protons and form molecular H2 [128, 129]. A fraction of the CO2 is

used as cell components and biomass, and the rest is released from the cells along with H2.

Another important function of ECH enzyme is the transfer of protons or sodium ions through the

cellular membrane creating an electrochemical proton gradient. This mechanism generates energy

through the ATP-synthesis that is initiated by the ATP-synthase [128, 129].

Figure 4.1 Schematic illustration of CO-oxidation, electron transfer, H2-formation, and proton transfer

through the cell-membrane. CODH: CO dehydrogenase; Fd: ferredoxin; ECH: energy-conserving

hydrogenase (adapted from [26, 128])

Page 27: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

34

34

During syngas fermentation, CO can be converted into CO2 (Eq. 3.1), and acetate can be

produced by acetogens via the reductive Acetyl-CoA or Wood-Ljungdahl pathway (Eq. 3.2 and

Eq. 3.4) as seen in Figure 4.2. The reductive Acetyl-CoA is an important pathway utilized by

microorganisms in syngas fermentation in order to achieve autotrophic CO2 fixation. This means

that carbon from an inorganic source is used by the cells as the sole carbon and energy source. In

syngas fermentation, several microorganisms follow the Acetyl-CoA pathway [68]. The Acetyl-

CoA synthase/CO-dehydrogenase complex (ACS/CODH) binds to a carbonyl and a methyl group

[130, 131], while CO-dehydrogenase enzyme reduces the CO2 to CO, which serves as a carbonyl

group [26]. The methyl group is formed after successive reductions of CO2 to several

intermediates, which are bound to a pterin coenzyme. The intermediate product Acetyl-CoA is

used as a cellular precursor and as an energy source. During the Acetyl-CoA pathway, H2 is the

electron donor in the presence of the hydrogenase enzyme. If H2 is absent, CO becomes the

electron donor, in the presence of CO-dehydrogenase (CODH). Therefore, the supply of adequate

H2 or CO during syngas fermentation is crucial [132]. Energy (1 mol ATP) and electrons are

required in order to reduce the CO2 to Acetyl-CoA, and the same amount of energy is recovered

when the acetyl-synthase catalyzes the Acetyl-CoA and CO reaction for acetate formation. In the

Acetyl-CoA pathway, 1 mol ATP is consumed while the CO2-pterin is reduced to methyl-pterin

[133], and 1 mol ATP is generated when acetate is formed. This energy conservation is done via

the substrate level phosphorylation.

35

Figure 4.2 The Wood-Ljungdahl pathway. H4F: tetrahydropholate; CoFeSP: corrinoid iron-sulphur protein

(adapted from [134]).

Acetate can be converted into CH4 and CO2 via aceticlastic methanogenesis as shown in

Figure 4.3. Initially, acetate is converted into acetyl-CoA, and then the methyl group is attached

to the tetrahydrosarcinapterin, and consequently to coenzyme M. Finally, the methyl-SCoM is

reduced to CH4 with coenzyme B as the electron donor [135]. In this pathway, 1 mol ATP is

generated for 1 mol of CH4 produced [136]. All acetotrophic methanogens gain energy by the

Page 28: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

36

36

electron transfer from ferredoxin to CoM-S-S-CoB with the creation of a proton gradient, as

shown in Figure 4.1 [135].

Figure 4.3 Metabolic pathway of aceticlastic methanogenesis. CoA: coenzyme A; H4SPT:

tetrahydrosarcinapterin; HS-CoM: sulfhydryl coenzyme M; HS-CoB: sulfhydryl coenzyme B; CoM-S-S-CoB:

heterodisulfide coenzymes CoM and CoB (adapted from [135]).

In hydrogenotrophic methanogenesis, the CO2 is converted into CH4 by the use of H2.

Therefore, this pathway controls the H2 to low levels and therefore assists the acetogenesis

reaction. Figure 4.4 shows the so-called Wolfe cycle of hydrogenotrophic methanogenesis [137].

The pathway has been described in literature [138] and starts with the reduction of CO2 to formyl

methanofuran with ferredoxin as the electron donor. Thereafter, the formyl group is attached to

the tetrahydromethanopterin (electron carrier) and then converted into methenyl-

tetrahydromethanopterin. Consequently, the electron donor reduces the electron carrier F420 to

F420H2 with hydrogenases and dehydrogenases. F420 catalyzes the reduction of methylene-

37

H4MPT to methyl-H4MPT. Afterwards, the methyl group is transported to the coenzyme M and

reduced to CH4. Energy is conserved by the electron bifurcation that occurs, heterodisulfide

reductase complex, when the CoM-S-S-CoB heterosulphide is converted into HS-CoM and HS-

CoB. This reaction generates the energy required for the formyl-MFR formation reaction [137,

139, 140]. Hydrogenotrophic methanogens can also use H2 in order to oxidize CO to CO2, which

is thereafter reduced to CH4 by utilizing H2 [141].

Figure 4.4 Metabolic pathway of hydrogenotrophic methanogenesis. MFR: methanofuran; H4MPT:

tetrahydromethanopterin; HS-CoM: sulfhydryl coenzyme M; HS-CoB: sulfhydryl coenzyme B; CoM-S-S-

CoB: heterodisulfide coenzymes CoM and CoB (adapted from [138]).

Page 29: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

38

38

4.4 Operational parameters

4.4.1 pH

The pH can affect the microbial activity during syngas fermentation. More specifically, it

can influence the cellular metabolism, by altering the intracellular pH and the electrochemical

gradient across the membrane [110]. The methanogens grow mostly at a neutral pH at a range of

6.8 8.5 [142]. Thus, most of the syngas fermentation studies are operated at pHs close to 7.0

[110]. The acetogenic bacteria are possibly the most tolerant group in various pH values,

including both alkaline and acidic environments [143]. These microorganisms are known to

produce acids at a higher pH and alcohols at a lower pH [144]. In addition, the hydrogenogens

mostly thrive at a neutral pH [145]. In a study on the mesophilic syngas biomethanation in

different pH and syngas pressures, the highest CH4 production of 0.89 mol CH4/mol syngas was

achieved for pH 5.8 and syngas pressure 1atm [146]. In this thesis, the effect of different pH

values (5, 6, and 7) was investigated during the addition of various concentrations of three heavy

metals (Paper III). According to the results, at high metal concentrations of 0.625 mg Cu/L, 3.75

mg Zn/L, and 17.5 mg Mn/L, the microbial activity was stimulated at pH 5 [66].

4.4.2 Temperature

The temperature in the bioreactors is of vital importance for the microbial activity in syngas

fermentation. The range of temperature affects the metabolic pathways used by the

microorganisms [110]. For example, acetate is considered the main CH4 precursor at mesophilic

conditions, while H2 is considered the main CH4 precursor in thermophilic conditions probably

because of the higher diversity of carboxydotrophic hydrogenogens in thermophilic conditions

[110, 147]. Another possible reason could be that the H2 formation from CO is

thermodynamically favored at thermophilic conditions [148]. On the other hand, even the

hydrogenotrophic methanogenesis is enhanced at thermophilic conditions in comparison to

mesophilic conditions. A study on the enrichment of mesophilic and thermophilic mixed culture

showed that CH4 productivity was approximately 18 times higher in mesophilic conditions and

that the thermophilic metabolism was dominated by carboxydotrophic hydrogenogens and

hydrogenotrophic methanogens [149]. Similarly, another study reported that the CO conversion

rates to CH4 were increased along with higher fermentation temperatures [147]. In this thesis,

(paper I), higher CH4 yields were observed during the conversion of artificial syngas (H2, CO,

39

CO2) by mixed culture, at 55 °C in comparison to 35 °C [150]. Another study reported that

various types of sludge had higher CO conversion to CH4, at thermophilic conditions, even if

they had been previously incubated at mesophilic conditions [147]. In general, higher syngas

conversion rates and CH4 production rates have been observed in thermophilic conditions;

however, higher energy input is also required and lower gas solubility can occur, in comparison

to mesophilic fermentation. The coupling of gasification and syngas fermentation could be an

advantage due to the fact that less effort is needed for the cooling of the syngas to thermophilic

levels than to mesophilic. Moreover, the lower gas solubility at a higher temperature can be

balanced by the increased gas diffusivity. For example, a study on gas diffusivity in water

showed that the CO diffusivity doubled from 35 °C to 60 °C [151].

4.4.3 Pressure

The pressure can affect the mass transfer rates of syngas components and consequently the

syngas conversion rates. An increase in the bioreactor´s operating pressure can increase the gas

solubility and achieve higher conversion rates without the need for bigger bioreactors [106, 152,

153]. The conversion of H2/CO2 mixture into CH4 has been studied in different pressures. In a

study on the influence of pressure in trickle bed bioreactors, the CH4 content increased by 34%

by increasing the pressure from 1.5 to 9 bar [153]. However, based on the author´s knowledge,

there are no studies on the effect of pressure during syngas conversion to H2 and CH4.

Theoretically, high CO pressures can cause inhibitory effects on the microbial activity. The effect

of the partial pressure of syngas components is discussed below.

4.4.4 Partial pressure

The partial pressure of each syngas component affects the mass transfer of this component

from the gas to the liquid phase inside the bioreactor´s broth. H2 and CO have low solubility in

liquids; thus, their partial pressure is important for efficient syngas conversion. On the other

hand, high CO partial pressure can lead to cellular inhibition. During syngas fermentation, the

most sensitive microorganisms to CO are carboxydotrophic methanogens and sulphate reducers,

which can usually tolerate PCO of 0.5 1.0 atm and 0.2 0.5 atm, respectively [121, 122, 154]. The

anaerobic culture can be more tolerant to high PCO after adaptation. A study reported that

although methanogenesis was totally inhibited at PCO ≥1 atm, higher methanogenic potential was

obtained after acclimatization to a 100% CO atmosphere [155]. This adaptation favored the

Page 30: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

40

40

domination of hydrogenotrophic methanogens, which were able to take over acetoclastic

methanogens, whereas syntrophic acetate oxidizing (SAO) bacteria converted acetate into CO2

and H2. Another study also reported the shift in microbial community from the addition of CO

during H2/CO2 fermentation. More specifically, the addition of CO resulted in the enrichment of

the microbial community with bacteria from the genera Thermincola and Thermoanaerobacter

[156]. The PH2 is considered to have less negative effect, in comparison to PCO, on the microbial

community. A study reported that the hydrogenase activity improved, along with the PH2, until

inhibition occurred probably due to saturation of the hydrogenase [157]. Another work reported

that increasing the initial pressure of H2/CO2 from 1 to 5 bar led to an increase in the CH4 rate

production from 0.035 ± 0.014 mmol h-1 to 0.072 ± 0.019 mmol h-1 [158].

4.5 Fermentation systems for hydrogen and methane production

The aim of the different bioreactor designs is to address the limitations of the syngas

fermentation process. Various bioreactor types have different advantages and disadvantages. For

example, the continuous stirred tank bioreactor (CSTR) has high mass transfer rates although it

requires high energy input [101, 159]. Trickle bed bioreactors have a lower energy demand than

CSTR and good gas-to-cell contact; however, gas channeling is a potential challenge [96, 160].

Membrane bioreactors (MBR), such as hollow fiber BR, offer high mass transfer rates, although

there is the challenge of clogging due to microbial growth on the fibers [161].

4.5.1 Methane production

Several studies have investigated the production for CH4 via syngas fermentation. The

conversion of CO into CH4 was studied in a continuous CSTR at both mesophilic and

thermophilic conditions with a mixed culture [162]. The H2:CO2 was fed in the bioreactor with a

ratio of 4:1, according to the stoichiometric requirements for hydrogenotrophic methanogenesis

(Eq. 3.5), and the cells were recycled in order to achieve high cell density. According to the

results, the highest CH4 productivity of 446 mmol LR-1 h-1 was achieved at thermophilic

conditions. Syngas biomethanation, with different types of anaerobic sludge, was also studied in

an upflow anaerobic sludge bed (UASB) bioreactor, where a maximum methanogenic activity of

1 mmol CH4 g-1 VSS d-1 was obtained at a PCO of 0.2 atm [155]. Higher CO biomethanation rates

were reported in a CSTR, by the use of a thermophilic co-culture of Carboxydothermus

41

hydrogenoformans and Methanothermobacter thermoautotrophicus in comparison with the

Methanothermobacter thermoautotrophicus monoculture [163]. The maximum CO conversion

efficiency and CH4 production rate achieved in that study were 93% and 17.6 mmol LR-1 d-1,

respectively.

A mixed culture was also used for the biomethanation of syngas. More specifically, syngas

was converted into CH4 by a triculture, consisting of the photosynthetic bacterium R. rubrumand

the methanogens M. formicicum, and M. barkeri [164]. The bacterium was chosen for the

conversion of CO into CO2 and H2 by the water gas shift reaction and the methanogens for the

conversion of CO2 and H2 into CH4. Two types of bioreactors were utilized, a packed bubble

column and a trickle bed bioreactor. The bubble column bioreactor was filled with glass Raching

rings (6mm × 6 mm) and the liquid medium was fed at the top of the bioreactor, whereas the

syngas was fed at the bottom of the vessel. The trickle bed bioreactor had a column height of 515

mm, with an inner diameter of 51 mm. The packing material was 6.35-mm Intalox saddles; the

total volume of the packed section was 1052 mL, and the gas and liquid flowed downward

through the packing material. The results showed CO conversion rates of 100% and 79% in the

trickle bed and bubble column bioreactor, respectively and mass transfer coefficients of 780 h-1

and 3.5 h-1 in the trickle bed and the bubble column bioreactor, respectively. The same triculture

was later used for the comparison of syngas fermentation in 5-cm and 16.5-cm diameter trickle

bed bioreactors [165]. The results showed a yield of 0.20 mol CH4 mol-1 H2, which is comparable

to the theoretical yield of 0.20 mol CH4 mol-1 H2. However, the larger bioreactor showed poor

performance probably due to low cell density or because of poor distribution of the liquid through

the column.

A new type of anaerobic trickle bed bioreactor was used for the biomethanation of H2 and

CO2 [160]. The 61-L pilot scale bioreactor was operated as a plug flow, at mesophilic conditions

with immobilized culture on a biofilm without gas-recirculation. The results showed an

impressive 98% CH4 concentration and a CH4 productivity of 2.77 mmol LR-1 h-1 at a gas

retention time of 4 h. A 30 L gas-lift bioreactor was used in another work for the biomethanation

of syngas with granular sludge [166]. The industrial granular sludge was first used for mesophilic

batch carboxydotrophic methanogenesis; thereafter, it was introduced in a gas-lift bioreactor and

fed with CO. The highest CO-conversion efficiency achieved was 75% at a PCO of 0.6 atm and a

Page 31: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

42

42

gas recirculation ratio of 20:1. A 5-fold higher CO-conversion potential was achieved at

thermophilic conditions. The highest CH4 specific productivity obtained was 0.126 mmol·g-

1VSS·h-1 (95% of the theoretical CH4 yield), at a gas recirculation rate of 600 mL min-1 and gas

retention time of 8.6 d. In a similar work, a stirred tank bioreactor with 400 mL working volume

was operated at 55 °C, pH 7 for the CO biomethanation with sewage sludge as the inoculum

[167]. The gaseous substrate was fed to the bioreactor via a hollow fiber membrane (HFM), and a

high CH4 productivity of 3.7 mmol·LR-1·h-1was reported although the CH4 percentage in the

gaseous product was only 19.2%. This was probably a result of the poor CO solubility in the

medium and the high CO flow rate of 6.6 mL·min-1·LR-1. The main archaeal species involved in

the CO conversion were Methanosarcina barkeri and Methanothermobacter thermautotrophicus,

while the bacterial species could not be identified.

The direct biomethanation of H2 and CO2 was achieved in an anaerobic bioreactor, with

Methanobacterium thermoautotrophicum cells immobilized on a cellulose acetate membrane or

inside a porous silica-alumina ceramic support [168]. The ceramic bioreactor was cylindrical (30

mm × 70 mm), with an average pore size and porosity of 100 μ and of 79.7%, respectively. In the

membrane bioreactor, a CH4 production rate of 0.75 ml CH4 per cm2 contact area per h was

achieved, while the initial fixed-cell mass increased from 0.2 mg dry cell cm-2 of contact area to 1

mg cm-2 after 12 h of cultivation. On the other hand, in the ceramic bioreactor, a CH4 production

rate of 6 L CH4 Lceramic-1 LR

-1 was achieved while the methanogen growth was homogeneous

inside the ceramic up to 7 cm depth, with a cell density between 20 and 30 mg dry cell cm-3

ceramic.

In this thesis (paper II), a reverse membrane bioreactor was operated at thermophilic

conditions for the biomethanation of artificial syngas in continuous mode. The bioreactor was a

serum glass bottle with 600-mL working volume, closed with rubber caps. The inoculum was a

mixed culture from a food digester enclosed inside hydrophilic PVDF flat sheet membranes that

formed 6 cm × 3 cm rectangles by heat sealing. Each sachet contained 3 g of inoculum, and each

bioreactor contained 15 sachets, while it was filled with 475 mL of synthetic medium. Syngas

was fed at the bottom of the bioreactor, and it was continuously recirculated. Although this type

of bubble column bioreactor has moderate mass transfer rates, it demands lower power

consumption, maintenance, and operational costs in comparison to other bioreactor types [159,

43

169]. The results showed that the reverse membrane bioreactor retained successfully high cell

density by eliminating the cell washout for 154 days. The maximum CH4 productivity achieved

was 8.3 mmol·L-1·h-1.

4.5.2 Hydrogen production

The syngas bioconversion to H2 has not been investigated so extensively as the syngas

biomethanation. Nevertheless, several studies with various bioreactor designs have been

conducted. Two studies reported the use of a mesophilic continuous stirred tank reactor for the

production of H2 from syngas (20% H2, 15% Ar, 55% CO, and 10% CO2), with the anaerobic

photosynthetic bacterium Rhodopirillum rubrum [170, 171]. The highest H2 production activity

and productivity achieved were approximately 16 mmol gcell-1 h-1 and 24 mmol L-1 h-1,

respectively, at pH 6.5, agitation of 500 rpm, syngas flow rate of 14 mL min-1 and the use of

acetate for cell growth [170]. In another work, the isolated strain Rhodopseudomonas palustris

PT was used in batch mesophilic experiments for the conversion of CO-rich gas into H2 [172].

The maximum H2 production in response to the CO consumption (86%) and the highest H2

concentration of 33.5 mmol L-1 was achieved at a concentration of 1.5 g L-1 sodium acetate.

The performance of a hollow fiber membrane bioreactor was studied for the CO conversion

to H2, with Carboxydothermus hydrogenoformans [173]. The study reported maximum CO

conversion ratio and H2 production rate of 97.6% and 0.46 mol d-1 at a gas loading rate of 0.22

mol d-1 and a liquid recirculation speed of 1.5 L min-1. The same bioreactor and strain were used

for the bioconversion of CO into H2 under various operational conditions [174]. In this work, a

biofilm was formed on the hollow membrane surface inside the liquid medium. This allowed for

the highest CO conversion activity of 0.44 mol CO g VSS d-1 and a maximum H2 productivity of

125 mmol L-1 h-1, at 70°C, with a PCO of 2 atm and liquid velocity of 65 m·h−1. The same

bioconversion was operated by the same pure strain in an 35 L gas-lift bioreactor, with

continuous gas supply for three months, at 70 °C [117]. The results showed highest conversion

efficiency of up to 90%, with a CO conversion rate of 2.7 mol g-1VSS d-1 and a H2 productivity of

6.7 mmol L-1 d-1. This performance was limited due to low cell density of the bioreactor.

A continuous stirred bioreactor with a working volume of 5 L was operated at mesophilic

conditions for the CO conversion to H2 [175]. At the beginning of this experiment, sucrose was

Page 32: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

44

44

supplied to the cells as a carbon and energy source and thereafter, the substrate was switched to

CO under anaerobic conditions. The maximum H2 productivity obtained was 41 mmol L-1 d-1 and

the CO conversion efficiency was 61%, at gas retention of 5 min, pH 7, and agitation of 700 rpm.

Another study employed a similar strategy of sucrose feeding before switching to CO feeding in a

CSTR with a working volume of 3 L [118]. The innovation of this study was the use of an

isolated bacterium strain, Citrobacter sp. Y19, which grows on aerobically organic carbon source

and converts CO and water into H2 in anaerobic environments. The study reported a maximum

H2 productivity of 5.7 mmol L-1 d-1, at pH 7, 30 °C and 500 rpm agitation, and three times higher

H2 production activity (27.1 mmol gcell-1 h-1) in comparison to that of Rhodospirillum rubrum.

The use of a genetically modified Thermococcus onnurineus NA1 for CO-dependent H2

production was investigated in a CSTR, with a working volume of 2 L, at 80 °C, pH 6.1–6.2, and

agitation of 300 rpm [123]. The mutant strain MC01 had a 30-fold higher transcription of the

mRNA encoding CODH and hydrogenase. This resulted in a maximum H2 productivity of 123.5

mmol L-1 d-1 that was 3.8-fold higher than that of the wild strain. The catabolic activity of the

mixed culture can be affected by components that favor the H2-formation while they inhibit

methanogenesis. For example, during a study on the effects of heavy metals in syngas

fermentation with a mixed culture (paper III), the addition of HNO3 inhibited the production of

CH4 and favored the H2 production [66]. The experiments were conducted in repeated batch

conditions with serum glass bottle bioreactors, at 55 °C. The results of this work are further

discussed in the following chapter.

4.6 The reverse membrane bioreactor

The concept of reverse membrane bioreactors (RMBR) is relatively new, and it has been

initiated and studied in the University of Borås since 2012 [176, 177]. In order to prepare the

membrane sachets, a flat sheet polyvinylidine difluoride (PVDF) membrane was cut into a

square, with dimensions 6 × 6 cm. Then, the membrane was folded into half, and the long side

and one of the smaller ends were heat-sealed, forming an envelope. Thereafter, the inoculum was

placed inside the envelope, and the last side was finally heat-sealed. The membrane sachet

containing the anaerobic cells was then immersed into the liquid medium of the bioreactor.

Liquid substrate passed though the membrane pores by passive diffusion, and gaseous products

exited the membrane when the required pressure was built inside the sachets. Figure 4.5 shows a

45

schematic illustration of the RMBR system that was operated continuously for 154 days (paper

II).

Figure 4.5 Schematic illustration of the reverse membrane bioreactor (RMBR) in the semi-continuous

biomethanation process of syngas and organic substances. (a) Digesting sludge encased in PVDF

membrane, (b) Organic and nutrient medium, (c) Syngas, (d) Peristaltic pump, (e) Gas outlet, (f) Warm

water bath, (g) Effluent, (h) Flow meter, (i) Data analysis [169].

Syngas fermentation for H2 and CH4 production has been investigated in several works, and

important progress has been achieved. However, there are main challenges that still need to be

overcome prior to the commercialization of the process. The main challenges, such as cell

washout, toxic effect of inhibitors on the microbial activity, and low gas-to-liquid mass transfer

rates, are discussed in the following chapter.

Page 33: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

46

46

47

Chapter 5

Improving syngas fermentation

Syngas fermentation for CH4 and H2 production has several advantages in comparison to

the catalytic conversion processes, as discussed in the previous chapter. However, this is a

relatively new technology; thus, there are still challenges that need to be addressed. For example,

the low gas-to-liquid mass transfer rates of syngas components and the systematic cell washout in

continuous processes in combination with the low growth rates of microorganisms, especially

methanogens, cause low fermentation yields. The presence of toxic syngas contaminants can

inhibit the microbial activity and lead to bioreactor failure. Furthermore, the unsustainable use of

gasification ashes, containing heavy metals, poses high health risks and leads to environmental

pollution. The main task of this thesis was to address the challenges mentioned above.

5.1 Cell washout

In continuous fermentation processes, the cell washout from bioreactors has been a

bottleneck that causes low cell density and lower capacity. Especially when CH4 is the desirable

product of fermentation, cell-washout is more important because of the low growth rate of

methanogens. For example, during a continuous syngas fermentation experiment with mixed

suspended culture, the CH4 production rate started to decrease after 40 days of fermentation

because of systematic cell-washout [169]. In another study, low yields were observed for cell-

dilution rates greater than 0.55 d-1 [159]. Different strategies have been investigated in order to

Page 34: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

48

48

minimize the cell-washout. Cell immobilization in tubes or packing material with cavities can

provide high cell-density, better gas-to-cell contact, and no need for mechanical agitation [102,

106]. The concept of immobilized sludge was investigated in a horizontal-flow anaerobic

immobilized sludge (HAIS) bioreactor [178]. In another work, a horizontal packed-bed

bioreactor, fed with glucose-based synthetic wastewater, was used for the conversion of low

contents of organic matter into organic acids and H2 [179]. However, the gas channeling caused

by the cell overgrowth on the tubes and cavities and the limiting column dimensions are

considered to be the main drawbacks of this method [102, 106]. In a study that aimed to address

the improvement of the biogas production rates, the anaerobic cells were encapsulated in natural

membranes that were made with alginate, using chitosan or Ca2+ as counter-ions, together with

the addition of carboxymethylcellulose (CMC) [176]. The results of that study showed that cell

encapsulation is a promising technique in order to achieve high cell density in bioreactors.

Figure 5.1 Elimination of cell washout in a RMBR

MBRs have been also used in order to increase the cell density and stop cell washout. Both

submerged and side stream membrane setups have been investigated. The submerged modules do

not require recirculation; however, the membrane area in contact with the liquid broth is limited,

and concentration polarization may occur [180]. Cell recycling in a continuous MBR allowed for

higher yields at higher dilution rates during syngas fermentation [181]. In another study, high

49

COD removal and CH4 yield with no VFA accumulation was achieved in an anaerobic biofilm

MBR for wastewater treatment [182]. The concept of the use of membranes in RMBR for high

cell density is schematically illustrated in Figure 5.1. The preparation and setup of the RMBR

was explained in section 4.5. This type of bioreactor has been successfully used in a previous

study for improving the biogas production rates during anaerobic digestion [176]. In another

similar set up, the cells were encased in multiple membrane horizontal layers, placed inside a

packed MBR [183]. However, the membrane encasement concept had never been used in syngas

fermentation.

In this thesis, a RMBR was used in order to address cell washout in syngas fermentation.

The membrane-encased cells were retained successfully inside the bioreactor during syngas

fermentation in both batch and continuous experiments [150, 169]. In the repeated batch

experiment (Paper I), the RMBR was successfully operated for the first time in syngas

biomethanation. Higher CH4 production was observed at 55 °C in comparison to 35 °C, and the

addition of a synthetic organic medium improved the CH4 production. In the case of the

continuous experiment (Paper II), a RMBR was successfully operated for 154 days without any

visual cell leaking from the membranes.

Figure 5.2 Comparison of the CH4 production rate in RMBR and FCBR, during the biomethanation of

syngas and organic substances [169]

Page 35: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

50

50

Figure 5.2 shows the comparison of CH4 production rate, in a RMBR and a free cell

bioreactor (FCBR) during continuous syngas fermentation. The addition of a liquid organic

medium after period III and the systematic cell washout caused a decrease in the CH4 production

in the FCBR, while the CH4 production in the RMBR was significantly higher. This result

showed that the high cell density was crucial for the higher CH4 production rate inside the RMBR

and that the membranes can be used for a long period without failing.

5.2 Inhibitors

In a syngas fermentation process, several contaminants may act as inhibitors of the

microbial activity, especially that of the methanogens. However, the reported effect of inhibitors

differs in the various studies due to the fact that the inhibition levels depend on several

parameters, such as the inhibitor concentration, the composition of the inoculum, the microbial

acclimatization to inhibitors as well as the environmental and operational conditions [184].

Moreover, synergistic and antagonistic effects between anaerobic microorganisms can influence

their tolerance in inhibiting components [184]. At low concentrations, some inhibitors may not

affect the microbial activity or they can be used as substrates. Two common types of

contaminants that are formed or released during the gasification process are tars and heavy

metals. Tars are formed during gasification, and they are contaminants of raw syngas. Syngas

contaminants need to be removed after gasification by costly processes. High concentration of

heavy metals are contained inside gasification ashes, which can leak into the environment. The

effect of three heavy metals, in syngas fermentation, was studied in this thesis. Furthermore, the

use of the RMBR in order to protect the anaerobic cells against high concentrations of two

common tars was investigated.

5.2.1 Effect of heavy metals in anaerobic digestion

Heavy metals are important for the metabolism of the anaerobic consortium because many

of them are part of enzymes that catalyze anaerobic reactions [185]. However, high

concentrations of heavy metals may lead to toxic conditions for the cells. In anaerobic digesters

fed with waste, such as municipal sewage and sludge and industrial waste, heavy metals can be

present in high concentrations. Heavy metals can disrupt the enzymatic activity and structure by

binding on enzyme prosthetic groups or by binding with thiol on proteins [186, 187]. Unlike

51

other components, the microbial community cannot degrade heavy metals, which can instead

accumulate to toxic concentrations [187] and lead to heavy metal toxicity and failure of the

anaerobic digester.

The stimulatory or inhibitory effect of heavy metals depends on the type of

microorganisms, the metal concentration in the digester, the chemical form of metals, and other

parameters, such as the pH and redox potential [188-190]. Studies that observed 50% inhibition

of methanogenesis reported that metal toxicity decreased in the order of Cu > Zn > Ni [190-192].

Another study on anaerobic digestion of sewage sludge reported that metal toxicity decreased in

the order of Cr > Ni > Cu > Zn [193]. In addition, the toxicity of some metals can be either

synergistic or antagonistic. More specifically, most mixed heavy metals, such as Cr–Cd–Pb, and

Zn–Cu–Ni were found to act synergistically [192], although Ahring and Westermann [194]

reported that the presence of Ni decreased the toxicity of Cd and Zn. A review of the effects of

heavy metals in anaerobic digestion reported that metals, such as Cu, Zn, Ni, Cd, Cr, and Pb have

been overwhelmingly reported to have inhibitory activity, depending on their concentrations

[195]. On the other side, the same review reported that metals, such as iron can have stimulatory

effect [195]. For example, a study reported that the conversion of acetate, propionate, and

methanol, by granulated sludge, was stimulated by the continuous addition of trace metals in the

influent [185]. Higher H2 yields were also achieved after the addition of Cu and Cr ions, during

anaerobic fermentation [35].

Figure 5.3 Utilization of heavy metals from gasification ash during syngas fermentation

Page 36: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

52

52

Figure 5.4 Relationships between H2 production activity, Ah (%) = (mol H2 mol-1 H2 control) 100 (a), H2

yield, YH = mol H2 mol-1 CO fed (b), initial medium pH and heavy metal concentrations (mg L-1) [66]

Heavy metals are found in large concentrations in the ash of the gasifier, at the end of the

gasification process. The ash is rich in heavy metals in various forms, such as metal oxides and

metal chlorides [67, 196, 197]. The use of these ashes in road and other constructions raise

53

significant concerns about the environment and public health because of leaching [198]. Heavy

metals are considered a threat because they are classified as possible carcinogens [199].

Therefore, the addition of heavy metals deriving from gasification ashes (Figure 5.3) could

reduce the footprint of gasification and improve the yields in syngas fermentation.

In this thesis (paper III), the effect of a metal mixture of Cu, Zn, and Mn as well as the

effect of pH were investigated during the syngas conversion into fermentative H2 [66]. The heavy

metals were mixed in various concentrations, and the results showed that the utilization of these

components improved the fermentative H2 production. Moreover, the beneficial and inhibitory

concentrations were reported. Figure 5.4 shows the H2 production activity in

(mmol/mmolcontrol) 100% and H2 yield, in mol H2 mol-1 CO fed, at different mixtures of Cu, Zn,

and Mn at a pH of 5, 6, and 7. According to the results, heavy metal concentrations of up to 0.1

mg Cu L-1, 0.67 mg Zn L-1 and 2.8 mg Mn L-1 were beneficial at all pH values. On the contrary, at

a higher heavy metal concentration of 0.625 mg Cu L-1, 3.75 mg Zn L-1 and 17.5 mg Mn L-1,

stimulation was observed only at pH 5. This shift in the preference of lower pH at higher metal

concentrations could be connected to the types of different thriving anaerobic strains in the

culture [66].

5.2.2 Protective effect of rMBR against syngas contaminants

Raw syngas can contain contaminants, such as particulate matter, tars, NOx, S, N2, and

alkali [70], which can inhibit the cellular performance during syngas fermentation [110]. Tar is

one of the most common types of contaminants; thus, it has been the main focus of several gas

cleaning technologies and studies [200]. Tars can be classified into: 1) >7 carbon ring

compounds; 2) heterocyclic, such as phenol; 3) light aromatic, such as toluene, styrene, and

xylene; 4) light polyaromatic, such as naphthalene and anthracene; and 5) heavy polyaromatic,

such as fluoroanthracene, pyrene, and chrysene [201]. Tars are condensable organic compounds

that are created mainly at lower temperatures during gasification. Toluene and naphthalene are

two common contaminants in raw syngas [202], and they can be toxic at high concentrations

[203-205]. The presence of tars and the tar content in raw syngas depends highly on the

gasification operating conditions and the type of feedstock. For example, benzene and toluene

can be found in syngas from gasification of switchgrass [206]. Although there is a high content of

tars in raw syngas, there is insufficient knowledge on the inhibitory effect of different syngas

Page 37: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

54

54

contaminants and on the minimum gas-cleaning requirements prior to fermentation [110]. From a

review in the literature, different inhibition levels of tars, such as toluene and naphthalene, are

reported in different studies [203, 207, 208]. The microbial tolerance toward these compounds

depends greatly on the composition of the microbial culture, as several bacteria can degrade

toluene and naphthalene [204, 209]. Microbial adaptation is also an important factor for the

toluene and naphthalene effect [203]. Furthermore, studies have shown that the presence of other

compounds can influence the effect of tar presence in digesters. For instance, a study reported

that the presence of toluene inhibited the anaerobic digestion, whereas the addition of toluene and

methanol did not cause any inhibition effect on the microbial activity [207]. Toluene and

naphthalene in raw syngas can be removed during costly gas cleaning processes as discussed in

Chapter 3. However, if this cleaning process could be avoided, the overall cost of the syngas

fermentation process could drop considerably.

Figure 5.5 Protective effect of membranes against syngas contaminants

An alternative and cost-effective solution to the syngas cleaning methods is to use raw

syngas as substrate and cover the cells in the bioreactor with a protective membrane. Figure 5.5

illustrates the concept of membrane protection against syngas contaminants. This concept is

based on the hydrophilicity of the membrane surface and the hydrophobic nature of syngas

55

contaminants. Similar MBRs have been successfully used in a previous study that showed their

protective effect against D-limonene, an inhibitor from citrus pills [210].

Figure 5.6 The effect of naphthalene on the (A) CH4 production rate, mmol CH4 L-1 d-1; (B) CH4 production

activity, (mol mol-1control) 100%; and (C) the naphthalene concentration, g L-1; in the FCBR and the RMBR

[submitted manuscript]

Page 38: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

56

56

In this thesis (paper IV), a RMBR was used in order to protect the encased cells during

their exposure to high toluene and naphthalene concentrations. The experiment was operated in

continuous mode, at 55 °C while the bioreactors were continuously shaken inside a shaking water

batch. According to the results, the CH4 production rate was higher in the RMBR in comparison

to the FCBR. More specifically, at the highest toluene and naphthalene concentration of 6.44 g L-

1 and 2.47 g L-1, the CH4 production rate in the RMBR was approximately 92% and 15%,

respectively, higher than in the FCBR. Figure 5.6 shows the effect of naphthalene´s increasing

concentration on the CH4 production rate, in mmol CH4 L-1 d-1 and CH4 production activity in

(mol mol-1control) 100% in the FCBR and in the RMBR. It can be concluded that CH4 production

was significantly higher in the RMBR than in the FCBR, from day 81 until day 120 for a

naphthalene concentration of 1.8 g L-1. A similar effect was observed from the increasing

concentration of toluene.

5.3 Gas-to-liquid mass transfer

The low gas-to-liquid mass transfer of syngas components and especially of CO and H2 is a

main challenge toward a more efficient syngas fermentation process. Several approaches such as

different bioreactor design, feeding method for syngas, cell immobilization, and bioreactor

conditions have been investigated. Studies reported that the utilization of packed bed and MBRs

led to higher mass transfer rates in comparison to CSTRs and bubble column bioreactors [94].

Furthermore, the syngas feeding with microspargers and hollow fibers increased the gas-to-liquid

mass transfer rates [96] and created better gas-to-cell contact through the biofilm attached on the

hollow fibers [161]. Packed material has also been used in order to increase the mass transfer

rates of syngas for the production of biogas. For example, an anaerobic trickle bed bioreactor

filled with packed material achieved a high CH4 concentration of 98%, from H2 and CO2

conversion [160].

During syngas fermentation, the gas components are consumed by microorganisms;

therefore, the rate of gas-to-liquid mass transfer is crucial for the process. The mass transfer rate

is often evaluated by the calculation of the mass transfer coefficient. Syngas components face

several resistances inside the bioreactor until they reach the cells (Figure 5.7-a). In the case of

cell encasement inside the membranes, there is a potential extra mass transfer resistance, which

57

can be considered negligible, however [150]. The main mass transfer resistance of syngas

compounds is considered to be the liquid film surrounding the cells (Figure 5.7-b), while all

other resistances can be considered as negligible [211].

Figure 5.7 (a) Main resistances during the mass transfer of syngas from the gas phase to the site of

reaction in the cells. Movement 1) in the gas bubble, 2) across the gas-liquid interfacial, 3) through the

liquid film surrounding the gas bubble, 4) in the liquid bulk, 5) through the membrane pores, 6) inside

the membrane, 7) across the liquid film surrounding the microbial cell, 8) through the cell membrane, 9)

through the cell, and ending up in the site of the reaction. (b) Movement of A through the interfacial

boundary. CAL: concentration of A in the liquid phase; CAli concentration of A in the liquid boundary; CAGi

concentration of A in the gas boundary; and CAG concentration of A in the gas phase [211]

The rate of mass transfer of component A though the gas (NAG) and the liquid (NAL)

boundary phase (in gmol m-3 s-1) can be calculated by the following equations:

NAG = kG α (CAG – CAGi) (5.1)

57

can be considered negligible, however [150]. The main mass transfer resistance of syngas

compounds is considered to be the liquid film surrounding the cells (Figure 5.7-b), while all

other resistances can be considered as negligible [211].

Figure 5.7 (a) Main resistances during the mass transfer of syngas from the gas phase to the site of

reaction in the cells. Movement 1) in the gas bubble, 2) across the gas-liquid interfacial, 3) through the

liquid film surrounding the gas bubble, 4) in the liquid bulk, 5) through the membrane pores, 6) inside

the membrane, 7) across the liquid film surrounding the microbial cell, 8) through the cell membrane, 9)

through the cell, and ending up in the site of the reaction. (b) Movement of A through the interfacial

boundary. CAL: concentration of A in the liquid phase; CAli concentration of A in the liquid boundary; CAGi

concentration of A in the gas boundary; and CAG concentration of A in the gas phase [211]

The rate of mass transfer of component A though the gas (NAG) and the liquid (NAL)

boundary phase (in gmol m-3 s-1) can be calculated by the following equations:

NAG = kG α (CAG – CAGi) (5.1)

57

can be considered negligible, however [150]. The main mass transfer resistance of syngas

compounds is considered to be the liquid film surrounding the cells (Figure 5.7-b), while all

other resistances can be considered as negligible [211].

Figure 5.7 (a) Main resistances during the mass transfer of syngas from the gas phase to the site of

reaction in the cells. Movement 1) in the gas bubble, 2) across the gas-liquid interfacial, 3) through the

liquid film surrounding the gas bubble, 4) in the liquid bulk, 5) through the membrane pores, 6) inside

the membrane, 7) across the liquid film surrounding the microbial cell, 8) through the cell membrane, 9)

through the cell, and ending up in the site of the reaction. (b) Movement of A through the interfacial

boundary. CAL: concentration of A in the liquid phase; CAli concentration of A in the liquid boundary; CAGi

concentration of A in the gas boundary; and CAG concentration of A in the gas phase [211]

The rate of mass transfer of component A though the gas (NAG) and the liquid (NAL)

boundary phase (in gmolm-3s-1) can be calculated by the following equations:

NAG = kGα(CAG – CAGi) (5.1)

Page 39: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

58

58

NAL = kL α (CALi – CAL) (5.2)

Where kG and kL is the gas-phase and liquid-phase mass transfer coefficient (m s-1),

respectively; and α is the total interfacial area for the mass transfer (m-1); CAL is the concentration

of A in the liquid bulk; and CAli is the concentration of A in the liquid boundary; CAG is the

concentration of A in the gas phase; and CAGi is the concentration of A in the gas boundary

(gmol m-3).

The Eq. 5.1 and Eq. 5.2 can be simplified to Eq. 5.3 and Eq. 5.4, respectively [211].

NAG = kG α (CAG – C*AG) (5.3)

NAL = kL α (C*AL – CAL) (5.4)

Where mCAL = C*AG is the gas-phase concentration of A in equilibrium with CAL; CAG/m =

C*AL is the liquid-phase concentration of A in equilibrium with CAG, and m is the distribution

factor. The last equations are when the main mass transfer resistance is the gas-phase film

resistance or the liquid-phase film resistance. Therefore, the overall mass transfer coefficients

KGα and KLα can be replaced by kGα and kLα [211]. In syngas fermentation, H2 and CO have low

solubility in the aqueous solution; thus, the kGα is significantly larger than kLα, and Eq. 5.3 is the

main mass transfer equation of the system.

Higher mas transfer rates can be achieved with higher gas holdup. In this thesis (paper V),

a packed floating MBR was used (Figure 5.8) in order to increase the gas holdup of syngas

bubbles. The membrane barrier in the floating MBR blocked the free anodic movement of the

syngas bubbles that rose from the bottom of the bioreactor. This caused an increase in the gas

holdup and the gas-to-cell contact, which consequently resulted in higher syngas mass transfer

and conversion rates. The FCBR had a poor gas-to-cell contact because of the distribution of the

cells that were mainly concentrated at the bottom of the FCBR. According to the results [212],

the H2 and CO conversion rates in the floating MBR were approximately 38% and 28% higher in

comparison to the FCBR. Moreover, the increase in the thickness of the membrane bed led to

even higher syngas conversion rates.

59

Figure 5.8 Improvement of gas-to-liquid mass transfer rates in a floating MBR

Figure 5.9 shows a comparison of the H2 and CO consumption rates and CH4 and CO2

production rates in a floating MBR, a bioreactor containing floating membranes and free cells

(floating MBR/FCBR), a bioreactor with free cells growing on packed material (PBR) and a

FCBR. According to the results, the highest consumption and production rates were observed in

the floating MBR. More specifically, the highest H2 and CO consumption rates obtained were

18.47 and 39.67 mmol L-1 d-1, respectively. In the same bioreactor, the highest CH4 production

and the highest CH4 yield of 21.33 mmol CH4 L-1 d-1 and 0.36 mol CH4 mol-1 (H2 + CO),

respectively, were reported.

Page 40: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

60

60

Figure 5.9 Syngas biomethanation in floating MBR, floating MBR/FCBR, PBR, and FCBR during continuous

syngas feeding. Consumption of (a) H2 and (b) CO and production of (c) CO2 and (d) CH4 in mmol·L-1·d-1)

[212]

The results from this chapter showed that the efficiency of syngas fermentation for CH4 and

H2 production was improved by addressing the main challenges of the process. The use and

development of the rMBR technology was a main tool in this research. This work focused on a

part of the challenges and their parameters in syngas fermentation. It is clear that more research is

needed in order to achieve a holistic approach and address all the main bottlenecks. In the

following chapter, the main conclusions are presented, and future work is suggested.

61

Chapter 6

Conclusions and future work

6.1 Major experimental results and conclusions

Syngas fermentation is a flexible process with several advantages in comparison to other

processes for waste treatment. However, there are still challenges, such as the cell washout, the

high heavy metal content inside the remaining ashes after gasification, the toxic effect of syngas

impurities on the cells, and the low gas-to-liquid mass transfer rates that have to be tackled. This

work aimed at the improvement of syngas fermentation by focusing on the above challenges. The

main results from the experimental work are the following:

1. The cell washout was successfully eliminated by the membrane encasement in a RMBR.

The membrane system was still efficient after 154 days of syngas biomethanation.

2. The syngas conversion rates into H2 were improved by the addition of specific amounts of

heavy metals.

3. The use of a RMBR had a protective effect on the cells against high concentrations of

toluene and naphthalene.

4. The use of a floating MBR increased the gas holdup and thus the syngas conversion rates

in syngas fermentation.

Page 41: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

62

62

6.2 Future work

This work addressed the use of RMBR for the improvement of syngas fermentation by

focusing on specific challenges mentioned in the previous section. However, several other factors

still need to be investigated.

6.2.1 Membrane material

One important consideration in MBRs is the cost and durability of membranes. A

commercial PVDF membrane was utilized in this thesis. The cost of this membrane is considered

high for large-scale MBRs. However, in our lab, a composite membrane that contains textile

waste was also prepared. The results that are not presented in this thesis showed similar syngas

conversion rates with both types of membranes (composite and PVDF), which indicates that

membranes from reused material with lower cost can be efficient alternatives to commercial

membranes. More research on alternative materials, based on sustainability, cost, and durability

should be conducted in the future.

6.2.2 Bioreactor´s design

One of the most important factors in syngas fermentation is the bioreactor´s design due to

the limited mass transfer rates of syngas components in the liquid medium. The high efficiency of

the floating MBR showed that the RMBR can be improved by internal adjustments. Several

features such as the height to diameter ratio and the use of a microsparger should be studied for

the further improvement of the rMBR.

6.2.3 Effect of syngas contaminants and heavy metals

This work focused on the effect of specific syngas contaminants and heavy metals

contained in the gasification ash, during syngas fermentation. However, a wider range of

contaminants and heavy metals as well as their antagonistic and synergistic effects should be

investigated in order to reach higher efficiencies in syngas fermentation.

63

References

1. Show, K.Y., et al., Biohydrogen production: Current perspectives and the way forward. International Journal of Hydrogen Energy, 2012. 37(20): p. 15616-15631.

2. Chandolias, K., S. Pardaev, and M.J. Taherzadeh, Biohydrogen and carboxylic acids production from wheat straw hydrolysate. Bioresource Technology, 2016. 216: p. 1093-1097.

3. Nielsen, C., et al., Food waste conversion to microbial polyhydroxyalkanoates. Microbial biotechnology, 2017. 10(6): p. 1338-1352.

4. Consonni, S., M. Giugliano, and M. Grosso, Alternative strategies for energy recovery from municipal solid waste Part A: Mass and energy balances. Waste Management, 2005. 25(2): p. 123-35.

5. Gershman, B., Gasification of Non-Recycled Plastics From Municipal Solid Waste In the United States. GBB Solid Waste Management Consultants, 2013.

6. Beneroso, D., et al., Comparing the composition of the synthesis-gas obtained from the pyrolysis of different organic residues for a potential use in the synthesis of bioplastics. Journal of Analytical and Applied Pyrolysis, 2015. 111: p. 55-63.

7. Bengelsdorf, F.R., et al., Chapter Four - Bacterial Anaerobic Synthesis Gas (Syngas) and CO2+H2 Fermentation, in Advances in Applied Microbiology, S. Sariaslani and G.M. Gadd, Editors. 2018, Academic Press. p. 143-221.

8. Drzyzga, O., et al., New challenges for syngas fermentation: towards production of biopolymers. Journal of Chemical Technology and Biotechnology, 2015. 90(10): p. 1735-1751.

9. EPA. Types of Anaerobic Digesters. 2019 [cited 2019; Available from: https://www.epa.gov/anaerobic-digestion/types-anaerobic-digesters.

10. Scarlat, N., J.-F. Dallemand, and F. Fahl, Biogas: Developments and perspectives in Europe. Renewable Energy, 2018. 129: p. 457-472.

11. Dagnall, S., J. Hill, and D. Pegg, Resource mapping and analysis of farm livestock manures—assessing the opportunities for biomass-to-energy schemes. Bioresource technology, 2000. 71(3): p. 225-234.

12. Møller, H.B., S.G. Sommer, and B.K. Ahring, Methane productivity of manure, straw and solid fractions of manure. Biomass and bioenergy, 2004. 26(5): p. 485-495.

13. Thompson, E., Q. Wang, and M. Li, Anaerobic digester systems (ADS) for multiple dairy farms: A GIS analysis for optimal site selection. Energy policy, 2013. 61: p. 114-124.

14. Raboni, M., P. Viotti, and A.G. Capodaglio, A comprehensive analysis of the current and future role of biofuels for transport in the European Union (EU). Revista ambiente & agua, 2015. 10(1): p. 9-21.

15. WMW. Report: Global Biogas Market to Reach $50 billion by 2026 2017 [cited 2019; Available from: https://waste-management-world.com/a/report-global-biogas-market-to-reach-50-billion-by.

16. Research, G.V. Grand View Research, Inc: Global Acetic Acid Market Expected to Reach USD 12.20 billion by 2020. 2015 [cited 2019 5 June]; Available from: http://www.grandviewresearch.com/.

17. Mengistu, M.G., et al., A review on biogas technology and its contributions to sustainable rural livelihood in Ethiopia. Renewable and Sustainable Energy Reviews, 2015. 48: p. 306-316.

18. Kougias, P.G. and I. Angelidaki, Biogas and its opportunities—A review. Frontiers of Environmental Science & Engineering, 2018. 12(3): p. 14.

19. van Foreest, F., Perspectives for biogas in Europe. 2012: Oxford Institute for Energy Studies. 20. EBA, EBA Annual Report 2019. 2019.

Page 42: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

64

64

21. Xia, Z., Domestic biogas in a changing China: can biogas still meet the energy needs of China’s rural households. International Institute for Environmental and Development, London, 2013.

22. Dong, L. The progress of biomass energy and biogas in China. in 19th Scientific Energy Management and Innovation Seminar. 2012.

23. Markets, R.a. Global Hydrogen Generation Market 2017-2018 & 2026: Market Accounted for $103.20 Billion in 2017 and is Expected to Reach $207.48 Billion by 2026. 2018 [cited 2019; Available from: https://www.prnewswire.com/news-releases/global-hydrogen-generation-market-2017-2018--2026-market-accounted-for-103-20-billion-in-2017-and-is-expected-to-reach-207-48-billion-by-2026--300763535.html.

24. Balat, M., Potential importance of hydrogen as a future solution to environmental and transportation problems. International journal of hydrogen energy, 2008. 33(15): p. 4013-4029.

25. Bakhtyari, A., M.A. Makarem, and M.R. Rahimpour, Hydrogen Production Through Pyrolysis, in Encyclopedia of Sustainability Science and Technology, R.A. Meyers, Editor. 2017, Springer New York: New York, NY. p. 1-28.

26. Henstra, A.M., et al., Microbiology of synthesis gas fermentation for biofuel production. Current Opinion in Biotechnology, 2007. 18(3): p. 200-206.

27. Tufa, R.A., et al., Salinity gradient power-reverse electrodialysis and alkaline polymer electrolyte water electrolysis for hydrogen production. Journal of Membrane Science, 2016. 514: p. 155-164.

28. Forsberg, C.W., Future hydrogen markets for large-scale hydrogen production systems. International Journal of Hydrogen Energy, 2007. 32(4): p. 431-439.

29. Cherry, R.S., A hydrogen utopia? International Journal of Hydrogen Energy, 2004. 29(2): p. 125-129.

30. Hoang, D.L. and S.H. Chan, Modeling of a catalytic autothermal methane reformer for fuel cell applications. Applied Catalysis A: General, 2004. 268(1-2): p. 207-216.

31. Al-Baghdadi, M.A.S., Hydrogen–ethanol blending as an alternative fuel of spark ignition engines. Renewable Energy, 2003. 28(9): p. 1471-1478.

32. Fan, J.M., et al., Comparative exergy analysis of chemical looping combustion thermally coupled and conventional steam methane reforming for hydrogen production. Journal of Cleaner Production, 2016. 131: p. 247-258.

33. Skjånes, K., et al., Design and construction of a photobioreactor for hydrogen production, including status in the field. Journal of Applied Phycology, 2016. 28: p. 2205-2223.

34. Levin, D.B., L. Pitt, and M. Love, Biohydrogen production: prospects and limitations to practical application. International Journal of Hydrogen Energy, 2004. 29(2): p. 173-185.

35. Lin, C.-Y. and S.-H. Shei, Heavy metal effects on fermentative hydrogen production using natural mixed microflora. International Journal of Hydrogen Energy, 2008. 33(2): p. 587-593.

36. Gujer, W. and A.J.B. Zehnder, Conversion processes in anaerobic digestion. Water science and technology, 1983. 15(8-9): p. 127-167.

37. Jugnia, L.B., et al., Production and Consumption of Methane in Soil, Peat, and Sediments from a Hydro-Electric Reservoir (Robert-Bourassa) and Lakes in the Canadian Taiga, in Greenhouse Gas Emissions — Fluxes and Processes: Hydroelectric Reservoirs and Natural Environments, A. Tremblay, et al., Editors. 2005, Springer Berlin Heidelberg: Berlin, Heidelberg. p. 441-465.

38. Hedlund, B.P., et al., An integrated study reveals diverse methanogens, Thaumarchaeota, and yet-uncultivated archaeal lineages in Armenian hot springs. Antonie Van Leeuwenhoek, 2013. 104(1): p. 71-82.

39. Moore, T.R. and R. Knowles, Methane Emissions from Fen, Bog and Swamp Peatlands in Quebec. Biogeochemistry, 1990. 11(1): p. 45-61.

40. Mao, C., et al., Process performance and methane production optimizing of anaerobic co-digestion of swine manure and corn straw. Scientific reports, 2017. 7(1): p. 9379-9379.

65

41. Ali Shah, F., et al., Microbial ecology of anaerobic digesters: the key players of anaerobiosis. The Scientific World Journal, 2014. 2014: p. 183752-183752.

42. Vavilin, V.A., et al., Hydrolysis kinetics in anaerobic degradation of particulate organic material: An overview. Waste Management, 2008. 28(6): p. 939-951.

43. Pavlostathis, S.G. and E. Giraldo-Gomez, Kinetics of Anaerobic Treatment. Water Science and Technology, 1991. 24(8): p. 35-59.

44. Doi, R.H., et al., Cellulosomes from Mesophilic Bacteria. Journal of Bacteriology, 2003. 185(20): p. 5907.

45. Bae, J., et al., Cellulosome complexes: natural biocatalysts as arming microcompartments of enzymes. Journal of Molecular Microbiology and Biotechnology, 2013. 23(4-5): p. 370-8.

46. Schink, B., Energetics of syntrophic cooperation in methanogenic degradation. Microbiology and Molecular Biology Reviews 1997. 61(2): p. 262-80.

47. Gerardi, M.H., The microbiology of anaerobic digesters. 2003, Hoboken, N.J: John Wiley. 48. Giusti, L., A review of waste management practices and their impact on human health. Waste

Management, 2009. 29(8): p. 2227-2239. 49. Schnurer, A. and A. Jarvis, Microbiological handbook for biogas plants. Swedish Waste

Management U, 2010. 2009: p. 1-74. 50. Forgács, G., Biogas production from citrus wastes and chicken feather: pretreatment and co-

digestion. 2012, Chalmers University of Technology. 51. Deublein, D. and A. Steinhauser, Biogas from waste and renewable resources: an introduction.

2011: John Wiley & Sons. 52. Li, W., et al., Methane production through anaerobic digestion: Participation and digestion

characteristics of cellulose, hemicellulose and lignin. Applied Energy, 2018. 226: p. 1219-1228. 53. Kabir, M.M., Bioprocessing of Recalcitrant Substrates for Biogas Production, in Skrifter från

Högskolan i Borås. 2015, Högskolan i Borås: Borås. p. 65. 54. Wikandari, R., et al., Effect of Effluent Recirculation on Biogas Production Using Two-stage

Anaerobic Digestion of Citrus Waste. Molecules, 2018. 23(12). 55. Jameel, H. and D.R. Keshwani, Thermochemical conversion of biomass to power and fuels, in

Biomass to renewable energy processes. 2017, CRC Press. p. 375-422. 56. Psomopoulos, C.S., A. Bourka, and N.J. Themelis, Waste-to-energy: A review of the status and

benefits in USA. Waste Management, 2009. 29(5): p. 1718-1724. 57. Buekens, A. and K. Cen, Waste incineration, PVC, and dioxins. Journal of Material Cycles and

Waste Management 2011. 13(3): p. 190-197. 58. Arena, U., Process and technological aspects of municipal solid waste gasification. A review.

Waste Management, 2012. 32(4): p. 625-639. 59. Malkow, T., Novel and innovative pyrolysis and gasification technologies for energy efficient and

environmentally sound MSW disposal. Waste Management, 2004. 24(1): p. 53-79. 60. Boerrigter, H. and R. Rauch, Review of applications of gases from biomass gasification. ECN

Biomassa, Kolen en Milieuonderzoek, 2006. 20. 61. Rauch, R., J. Hrbek, and H. Hofbauer, Biomass gasification for synthesis gas production and

applications of the syngas. Wiley Interdisciplinary Reviews: Energy and Environment, 2014. 3(4): p. 343-362.

62. Knoef, H. and J. Ahrenfeldt, Handbook biomass gasification. 2005: BTG biomass technology group The Netherlands.

63. Klinghoffer, N., M.J. Castaldi, and A. Nzihou. Beneficial use of ash and char from biomass gasification. in 19th Annual North American Waste-to-Energy Conference. 2011: American Society of Mechanical Engineers.

Page 43: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

66

66

64. Fernández-Pereira, C., et al., Application of biomass gasification fly ash for brick manufacturing. Fuel, 2011. 90(1): p. 220-232.

65. Vestin, J., et al., Biofuel ash in road stabilization – Lessons learned from six years of field testing. Transportation Geotechnics, 2018. 14: p. 146-156.

66. Chandolias, K., Wainaina, S., Niklasson, C., Taherzadeh, M.J., Effects of heavy metals and pH on the conversion of biomass to hydrogen via syngas fermentation. Bioresources, 2018. 13(2): p. 4455-4469.

67. Tafur-Marinos, J.A., et al., Comparison of inorganic constituents in bottom and fly residues from pelletised wood pyro-gasification. Fuel, 2014. 119: p. 157-162.

68. Chandolias, K., T. Richards, and M.J. Taherzadeh, Chapter 5 - Combined Gasification-Fermentation Process in Waste Biorefinery, in Waste Biorefinery, T. Bhaskar, et al., Editors. 2018, Elsevier. p. 157-200.

69. Pfeifer, C., B. Puchner, and H. Hofbauer, Comparison of dual fluidized bed steam gasification of biomass with and without selective transport of CO2. Chemical Engineering Science, 2009. 64(23): p. 5073-5083.

70. Woolcock, P.J. and R.C. Brown, A review of cleaning technologies for biomass-derived syngas. Biomass and Bioenergy, 2013. 52: p. 54-84.

71. Babu, S.P., Thermal gasification of biomass technology developments: end of task report for 1992 to 1994. Biomass and Bioenergy, 1995. 9(1-5): p. 271-285.

72. van de Kamp, W.L., et al., Sampling and analysis of tar and particles in biomass producer gases. Energy Research Centre of the Netherlands, report ECN-C—06-046, 2005.

73. Carpenter, D., et al., Biomass feedstocks for renewable fuel production: a review of the impacts of feedstock and pretreatment on the yield and product distribution of fast pyrolysis bio-oils and vapors. Green Chemistry, 2014. 16(2): p. 384-406.

74. Abdoulmoumine, N., et al., A review on biomass gasification syngas cleanup. Applied Energy, 2015. 155: p. 294-307.

75. Hirohata, O., et al., Release behavior of tar and alkali and alkaline earth metals during biomass steam gasification. Energy and Fuels, 2008. 22(6): p. 4235-4239.

76. Sharma, S.D., et al., A critical review of syngas cleaning technologies — fundamental limitations and practical problems. Powder Technology, 2008. 180(1): p. 115-121.

77. Gautam, G., et al., Tar analysis in syngas derived from pelletized biomass in a commercial stratified downdraft gasifier. BioResources, 2011. 6(4): p. 4653-4661.

78. Hoekman, S.K., et al., Characterization of trace contaminants in syngas from the thermochemical conversion of biomass. Biomass Conversion and Biorefinery, 2013. 3(4): p. 271-282.

79. Tijmensen, M.J.A., et al., Exploration of the possibilities for production of Fischer Tropsch liquids and power via biomass gasification. Biomass and Bioenergy, 2002. 23(2): p. 129-152.

80. Hamelinck, C.N., et al., Production of FT transportation fuels from biomass; technical options, process analysis and optimisation, and development potential. Energy, 2004. 29(11): p. 1743-1771.

81. Boerrigter, H., H. Den Uil, and H.-P. Calis, Green diesel from biomass via Fischer-Tropsch synthesis: new insights in gas cleaning and process design. Pyrolysis and gasification of Biomass and waste, 2003. 1.

82. Leibold, H., A. Hornung, and H. Seifert, HTHP syngas cleaning concept of two stage biomass gasification for FT synthesis. Powder Technology, 2008. 180(1-2): p. 265-270.

83. Boerrigter, H., R. Rauch, and H.A.M. Knoef, Handbook Biomass Gasification. 2nd ed, ed. B.b.t. group. 2005, The Netherlands: Enschede.

84. Ogden, J., M. , Prospects for building a hydrogen energy infrastructure Annual Review of Energy and the Environment, 1999. 24(1): p. 227-279.

67

85. Ratnasamy, C. and J.P. Wagner, Water Gas Shift Catalysis. Catalysis Reviews, 2009. 51(3): p. 325-440.

86. Bouarab, R., et al., Hydrogen Production from the Water-Gas Shift Reaction on Iron Oxide Catalysts. Journal of Catalysts, 2014. 2014: p. 6.

87. Newsome, D.S., The Water - Gas Shift Reaction. Catalysis Reviews, 2006. 21(2): p. 275-318. 88. Chianese, S., et al., Hydrogen from the high temperature water gas shift reaction with an

industrial Fe/Cr catalyst using biomass gasification tar rich synthesis gas. Fuel Processing Technology, 2015. 132: p. 39-48.

89. Soukup, G., et al., In situ CO2 capture in a dual fluidized bed biomass steam gasifier–bed material and fuel variation. Chemical Engineering & Technology: Industrial Chemistry-Plant Equipment-Process Engineering-Biotechnology, 2009. 32(3): p. 348-354.

90. Kopyscinski, J., T.J. Schildhauer, and S.M.A. Biollaz, Production of synthetic natural gas (SNG) from coal and dry biomass – A technology review from 1950 to 2009. Fuel, 2010. 89(8): p. 1763-1783.

91. Liu, Z., et al., Total methanation of syngas to synthetic natural gas over Ni catalyst in a micro-channel reactor. Fuel, 2012. 95: p. 599-605.

92. Antonio, G.A., V.S. Ioannis, and N.G. Hariklia, Syngas biomethanation: state-of-the-art review and perspectives. Biofuels, Bioproducts and Biorefining, 2018. 12(1): p. 139-158.

93. Salman, C.A., et al., Enhancing biomethane production by integrating pyrolysis and anaerobic digestion processes. Applied Energy, 2017.

94. Asimakopoulos, K., H.N. Gavala, and I.V. Skiadas, Reactor systems for syngas fermentation processes: A review. Chemical Engineering Journal, 2018. 348: p. 732-744.

95. Luo, G. and I. Angelidaki, Integrated biogas upgrading and hydrogen utilization in an anaerobic reactor containing enriched hydrogenotrophic methanogenic culture. Biotechnology and Bioengineering, 2012. 109(11): p. 2729-36.

96. Bredwell, M.D., P. Srivastava, and R.M. Worden, Reactor Design Issues for Synthesis-Gas Fermentations. Biotechnology Progress, 1999. 15(5): p. 834-844.

97. Klasson, K.T., et al., Biological conversion of synthesis gas into fuels. International Journal of Hydrogen Energy, 1992. 17(4): p. 281-288.

98. Daniell, J., M. Köpke, and S. Simpson, Commercial Biomass Syngas Fermentation. Energies, 2012. 5(12): p. 5372.

99. Durre, P., Butanol formation from gaseous substrates. FEMS Microbiology Letters, 2016. 363(6). 100. Liew, F., et al., Gas Fermentation-A Flexible Platform for Commercial Scale Production of Low-

Carbon-Fuels and Chemicals from Waste and Renewable Feedstocks. Frontiers in Microbiology, 2016. 7: p. 694.

101. Munasinghe, P.C. and S.K. Khanal, Biomass-derived syngas fermentation into biofuels: Opportunities and challenges. Bioresource Technology, 2010. 101(13): p. 5013-5022.

102. Mohammadi, M., et al., Bioconversion of synthesis gas to second generation biofuels: A review. Renewable and Sustainable Energy Reviews, 2011. 15(9): p. 4255-4273.

103. Devarapalli, M. and H.K. Atiyeh, A review of conversion processes for bioethanol production with a focus on syngas fermentation. Biofuel Research Journal, 2015. 2(3): p. 268-280.

104. Xu, D., D.R. Tree, and R.S. Lewis, The effects of syngas impurities on syngas fermentation to liquid fuels. Biomass and Bioenergy, 2011. 35(7): p. 2690-2696.

105. Sipma, J., et al., Hydrogenogenic CO Conversion in a Moderately Thermophilic (55 °C) Sulfate-Fed Gas Lift Reactor: Competition for CO-Derived H2. Biotechnology Progress, 2006. 22(5): p. 1327-1334.

106. Klasson, K.T., et al., Bioreactor design for synthesis gas fermentations. Fuel, 1991. 70(5): p. 605-614.

Page 44: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

68

68

107. Lee, P.-H., et al., Enhancement of carbon monoxide mass transfer using an innovative external hollow fiber membrane (HFM) diffuser for syngas fermentation: Experimental studies and model development. Chemical Engineering Journal, 2012. 184(0): p. 268-277.

108. Shen, Y., R. Brown, and Z. Wen, Enhancing mass transfer and ethanol production in syngas fermentation of Clostridium carboxidivorans P7 through a monolithic biofilm reactor. Applied Energy, 2014. 136: p. 68-76.

109. Kleerebezem, R. and M.C.M. van Loosdrecht, Mixed culture biotechnology for bioenergy production. Current Opinion in Biotechnology, 2007. 18(3): p. 207-212.

110. Grimalt-Alemany, A., I.V. Skiadas, and H.N. Gavala, Syngas biomethanation: state-of-the-art review and perspectives. Biofuels, Bioproducts and Biorefining, 2017. 12(1): p. 139-158.

111. Hoogendoorn, A. and H.J. van Kasteren, Transportation biofuels: novel pathways for the production of ethanol. 2010: Royal Society of Chemistry.

112. Demler, M. and D. Weuster-Botz, Reaction engineering analysis of hydrogenotrophic production of acetic acid by Acetobacterium woodii. Biotechnology and Bioengineering, 2011. 108(2): p. 470-4.

113. Heiskanen, H., I. Virkajärvi, and L. Viikari, The effect of syngas composition on the growth and product formation of Butyribacterium methylotrophicum. Enzyme and Microbial Technology, 2007. 41(3): p. 362-367.

114. Sim, J.H., et al., Clostridium aceticum—A potential organism in catalyzing carbon monoxide to acetic acid: Application of response surface methodology. Enzyme and Microbial Technology, 2007. 40(5): p. 1234-1243.

115. Younesi, H., G. Najafpour, and A.R. Mohamed, Ethanol and acetate production from synthesis gas via fermentation processes using anaerobic bacterium, Clostridium ljungdahlii. Biochemical Engineering Journal, 2005. 27(2): p. 110-119.

116. Kundiyana, D.K., R.L. Huhnke, and M.R. Wilkins, Syngas fermentation in a 100-L pilot scale fermentor: design and process considerations. Journal of Bioscience and Bioengineering, 2010. 109(5): p. 492-498.

117. Haddad, M., R. Cimpoia, and S.R. Guiot, Performance of Carboxydothermus hydrogenoformans in a gas-lift reactor for syngas upgrading into hydrogen. International Journal of Hydrogen Energy, 2014. 39(6): p. 2543-2548.

118. Jung, G.Y., et al., Hydrogen production by a new chemoheterotrophic bacterium Citrobacter sp. Y19. International Journal of Hydrogen Energy, 2002. 27(6): p. 601-610.

119. Küsel, K., et al., Clostridium scatologenes strain SL1 isolated as an acetogenic bacterium from acidic sediments. International Journal of Systematic and Evolutionary Microbiology, 2000. 50(2): p. 537-546.

120. O'Brien, J.M., et al., Association of hydrogen metabolism with unitrophic or mixotrophic growth of Methanosarcina barkeri on carbon monoxide. Journal of Bacteriology, 1984. 158(1): p. 373-375.

121. Rother, M. and W.W. Metcalf, Anaerobic growth of Methanosarcina acetivorans C2A on carbon monoxide: An unusual way of life for a methanogenic archaeon. Proceedings of the National Academy of Sciences of the United States of America, 2004. 101(48): p. 16929-16934.

122. Daniels, L., et al., Carbon monoxide oxidation by methanogenic bacteria. Journal of Bacteriology, 1977. 132(1): p. 118-26.

123. Kim, M.S., et al., CO-dependent H2 production by genetically engineered Thermococcus onnurineus NA1. Applied Environmental Microbiology, 2013. 79(6): p. 2048-53.

124. Asimakopoulos, K., H.N. Gavala, and I.V. Skiadas, Biomethanation of Syngas by Enriched Mixed Anaerobic Consortia in Trickle Bed Reactors. Waste and Biomass Valorization, 2019.

69

125. Latif, H., et al., Trash to treasure: production of biofuels and commodity chemicals via syngas fermenting microorganisms. Current Opinion in Biotechnology, 2014. 27: p. 79-87.

126. Buckel, W. and R.K. Thauer, Energy conservation via electron bifurcating ferredoxin reduction and proton/Na+ translocating ferredoxin oxidation. Biochimica et Biophysica Acta (BBA) - Bioenergetics, 2013. 1827(2): p. 94-113.

127. Schuchmann, K. and V. Muller, Autotrophy at the thermodynamic limit of life: a model for energy conservation in acetogenic bacteria. Nature Reviews Microbiology, 2014. 12(12): p. 809-21.

128. Hedderich, R., Energy-Converting [NiFe] Hydrogenases from Archaea and Extremophiles: Ancestors of Complex I. Journal of Bioenergetics and Biomembranes, 2004. 36(1): p. 65-75.

129. Singer, S.W., M.B. Hirst, and P.W. Ludden, CO-dependent H2 evolution by Rhodospirillum rubrum: Role of CODH:CooF complex. Biochimica et Biophysica Acta - Bioenergetics, 2006. 1757(12): p. 1582-1591.

130. Ragsdale, S.W., Life with carbon monoxide. Critical Reviews in Biochemistry and Molecular Biology, 2004. 39(3): p. 165-195.

131. Lindahl, P.A., The Ni-containing carbon monoxide dehydrogenase family: Light at the end of the tunnel? Biochemistry, 2002. 41(7): p. 2097-2105.

132. Munasinghe, P., Chaminda, and S. Khanal, Kumar, Biomass-derived Syngas Fermentation into Biofuels, in Biofuels: Alternative Feedstocks and Conversion Processes, A. Pandey, et al., Editors. 2012, Elsevier. p. 79-98.

133. Wheelis, M., Principles of modern microbiology. 2011: Jones & Bartlett Publishers. 134. Nitschke, W. and M.J. Russell, Beating the acetyl coenzyme A-pathway to the origin of life.

Philosophical Transactions of the Royal Society of London B: Biological Sciences, 2013. 368(1622).

135. Ferry, J.G., Fundamentals of methanogenic pathways that are key to the biomethanation of complex biomass. Current Opinion in Biotechnology, 2011. 22(3): p. 351-357.

136. Deppenmeier, U. and V. Muller, Life close to the thermodynamic limit: how methanogenic archaea conserve energy. Results and Problems in Cell Differentiation, 2008. 45: p. 123-52.

137. Thauer, R.K., The Wolfe cycle comes full circle. Proceedings of the National Academy of Sciences, 2012. 109(38): p. 15084-15085.

138. Zabranska, J. and D. Pokorna, Bioconversion of carbon dioxide to methane using hydrogen and hydrogenotrophic methanogens. Biotechnology Advances, 2018. 36(3): p. 707-720.

139. Kaster, A.-K., et al., Coupling of ferredoxin and heterodisulfide reduction via electron bifurcation in hydrogenotrophic methanogenic archaea. Proceedings of the National Academy of Sciences, 2011. 108(7): p. 2981-2986.

140. Martin, W.F., Hydrogen, metals, bifurcating electrons, and proton gradients: the early evolution of biological energy conservation. FEBS letters, 2012. 586(5): p. 485-493.

141. Guneratnam, A.J., et al., Study of the performance of a thermophilic biological methanation system. Bioresource technology, 2017. 225: p. 308-315.

142. Liu, Y. and W.B. Whitman, Metabolic, phylogenetic, and ecological diversity of the methanogenic archaea. Annals of the New York Academy of Sciences, 2008. 1125: p. 171-89.

143. Drake, H.L., K. Küsel, and C. Matthies, Acetogenic prokaryotes. The Prokaryotes: Volume 2: Ecophysiology and Biochemistry, 2006: p. 354-420.

144. Grupe, H. and G. Gottschalk, Physiological Events in Clostridium acetobutylicum during the Shift from Acidogenesis to Solventogenesis in Continuous Culture and Presentation of a Model for Shift Induction. Applied and environmental microbiology, 1992. 58(12): p. 3896-3902.

145. Sipma, J., et al., Microbial CO conversions with applications in synthesis gas purification and bio-desulfurization. Critical Reviews in Biotechnology, 2006. 26(1): p. 41-65.

Page 45: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

70

70

146. Pereira, F.M.R., M.M. Alves, and D.Z. Sousa. Effect of pH and pressure on syngas fermentation by anaerobic mixed cultures. in 13th World Congress on Anaerobic Digestion. 2013.

147. Sipma, J., et al., Carbon monoxide conversion by anaerobic bioreactor sludges. FEMS Microbiology Ecology, 2003. 44(2): p. 271-277.

148. Conrad, R. and B. Wetter, Influence of temperature on energetics of hydrogen metabolism in homoacetogenic, methanogenic, and other anaerobic bacteria. Archives of Microbiology, 1990. 155(1): p. 94-98.

149. Grimalt-Alemany, A., et al., Enrichment of Mesophilic and Thermophilic Mixed Microbial Consortia for Syngas Biomethanation: The Role of Kinetic and Thermodynamic Competition. Waste and Biomass Valorization, 2019.

150. Youngsukkasem, S., K. Chandolias, and M. Taherzadeh, J., Rapid bio-methanation of syngas in a reverse membrane bioreactor: Membrane encased microorganisms. Bioresource Technology, 2015. 178: p. 334-340.

151. Wise, D.L. and G. Houghton, Diffusion coefficients of neon, krypton, xenon, carbon monoxide and nitric oxide in water at 10–60 C. Chemical Engineering Science, 1968. 23(10): p. 1211-1216.

152. Vega, J.L., E.C. Clausen, and J.L. Gaddy, Design of bioreactors for coal synthesis gas fermentations. Resources, Conservation and Recycling, 1990. 3(2–3): p. 149-160.

153. Ullrich, T., et al., Influence of operating pressure on the biological hydrogen methanation in trickle-bed reactors. Bioresource Technology, 2018. 247: p. 7-13.

154. Parshina, S.N., et al., Desulfotomaculum carboxydivorans sp. nov., a novel sulfate-reducing bacterium capable of growth at 100% CO. International Journal of Systematic and Evolutionary Microbiology, 2005. 55(5): p. 2159-2165.

155. Sancho Navarro, S., et al., Biomethanation of Syngas Using Anaerobic Sludge: Shift in the Catabolic Routes with the CO Partial Pressure Increase. Frontiers in Microbiology, 2016. 7: p. 1188.

156. Alves, J.I., et al., Enrichment of anaerobic syngas-converting bacteria from thermophilic bioreactor sludge. FEMS Microbiology Ecology, 2013. 86(3): p. 590-7.

157. Skidmore, B.E., et al., Syngas fermentation to biofuels: Effects of hydrogen partial pressure on hydrogenase efficiency. Biomass and Bioenergy, 2013. 55: p. 156-162.

158. Lopes, M., et al. Hydrogenotrophic activity under increased H2/CO2 pressure: Effect on methane production and microbial community. in Abstracts European Biotechnology Contress 2015. 2015.

159. Datar, R.P., et al., Fermentation of biomass-generated producer gas to ethanol. Biotechnology and Bioengineering 2004. 86: p. 587 - 594.

160. Burkhardt, M., T. Koschack, and G. Busch, Biocatalytic methanation of hydrogen and carbon dioxide in an anaerobic three-phase system. Bioresource Technology, 2015. 178: p. 330-333.

161. Shen, Y.W., R. Brown, and Z.Y. Wen, Syngas fermentation of Clostridium carboxidivoran P7 in a hollow fiber membrane biofilm reactor: Evaluating the mass transfer coefficient and ethanol production performance. Biochemical Engineering Journal, 2014. 85: p. 21-29.

162. Wise, D.L., C.L. Cooney, and D.C. Augenstein, Biomethanation: Anaerobic fermentation of CO2, H2 and CO to methane. Biotechnology and Bioengineering, 1978. 20(8): p. 1153-1172.

163. Diender, M., et al., High Rate Biomethanation of Carbon Monoxide-Rich Gases via a Thermophilic Synthetic Coculture. ACS Sustainable Chemistry & Engineering, 2018. 6(2): p. 2169-2176.

164. Klasson, K.T., et al., Methane production from synthesis gas using a mixed culture ofR. rubrum M. barkeri, and M. formicicum. Applied Biochemistry and Biotechnology, 1990. 24(1): p. 317-328.

165. Kimmel, D.E., et al., Performance of trickle-bed bioreactors for converting synthesis gas to methane. Applied Biochemistry and Biotechnology, 1991. 28-29: p. 457-69.

166. Guiot, S.R., R. Cimpoia, and G. Carayon, Potential of wastewater-treating anaerobic granules for biomethanation of synthesis gas. Environmental Science & Technology, 2011. 45(5): p. 2006-12.

71

167. Luo, G., W. Wang, and I. Angelidaki, Anaerobic Digestion for Simultaneous Sewage Sludge Treatment and CO Biomethanation: Process Performance and Microbial Ecology. environmental Science Technology, 2013. 47(18): p. 10685-10693.

168. Jee, H.S., et al., Biomethanation of H2 and CO2 by Methanobacterium thermoautotrophicum in membrane and ceramic bioreactors. Journal of Fermentation Technology, 1987. 65(4): p. 413-418.

169. Westman Youngsukkasem, S., K. Chandolias, and J.M. Taherzadeh, Syngas Biomethanation in a Semi-Continuous Reverse Membrane Bioreactor (RMBR). Fermentation, 2016. 2(8): p. 1-12.

170. Younesi, H., et al., Biohydrogen production in a continuous stirred tank bioreactor from synthesis gas by anaerobic photosynthetic bacterium: Rhodopirillum rubrum. Bioresource Technology, 2008. 99(7): p. 2612-9.

171. Klasson, K.T., et al., Evaluation of mass-transfer and kinetic parameters for Rhodospirillum rubrum in a continuous stirred tank reactor. Applied Biochemistry and Biotechnology, 1993. 39-40(1): p. 549-557.

172. Pakpour, F., et al., Biohydrogen production from CO-rich syngas via a locally isolated Rhodopseudomonas palustris PT. Bioprocess and Biosystems Engineering, 2014. 37(5): p. 923-30.

173. Zhao, Y., Z.J. Liu, and S.R. Guiot, Continuous bio-hydrogen production from syngas fermentation in a hollow fiber membrane reactor. Xiandai Huagong/Modern Chemical Industry, 2011. 31(9): p. 71-74+76.

174. Zhao, Y., et al., Performance of a Carboxydothermus hydrogenoformans-immobilizing membrane reactor for syngas upgrading into hydrogen. International Journal of Hydrogen Energy, 2013. 38(5): p. 2167-2175.

175. Oh, Y.-K., et al., Biohydrogen production from carbon monoxide and water byRhodopseudomonas palustris P4. Biotechnology and Bioprocess Engineering, 2005. 10(3): p. 270.

176. Youngsukkasem, S., S.K. Rakshit, and M.J. Taherzadeh, Biogas production by encapsulated methane-producing bacteria. Bioresources, 2012. 7(1): p. 56-65.

177. Mahboubi, A., et al., Reverse membrane bioreactor: Introduction to a new technology for biofuel production. Biotechnology Advances, 2016. 34(5): p. 954-975.

178. Zaiat, M., et al., Rational basis for designing horizontal-flow anaerobic immobilized sludge (HAIS) reactor for wastewater treatment. Brazilian Journal of Chemical Engineering, 1997. 14.

179. Leite, J.A.C., et al., Application of an anaerobic packed-bed bioreactor for the production of hydrogen and organic acids. International Journal of Hydrogen Energy, 2008. 33(2): p. 579-586.

180. Cui, Z.F. and H.S. Muralidhara, Membrane technology: a practical guide to membrane technology and applications in food and bioprocessing. 2010: Elsevier.

181. Kargupta, K., S. Datta, and S.K. Sanyal, Analysis of the performance of a continuous membrane bioreactor with cell recycling during ethanol fermentation. Biochemical Engineering Journal, 1998. 1(1): p. 31-37.

182. Li, N., et al., Robust performance of a novel anaerobic biofilm membrane bioreactor with mesh filter and carbon fiber (ABMBR) for low to high strength wastewater treatment. Chemical Engineering Journal, 2017. 313: p. 56-64.

183. Youngsukkasem, S., et al., Rapid Biogas Production by Compact Multi-Layer Membrane Bioreactor: Efficiency of Synthetic Polymeric Membranes. Energies, 2013. 6(12): p. 6211-6224.

184. Chen, Y., J.J. Cheng, and K.S. Creamer, Inhibition of anaerobic digestion process: a review. Bioresource Technology, 2008. 99(10): p. 4044-64.

185. Osuna, M., B., , et al., Effects of Trace Element Addition on Volatile Fatty Acid Conversions in Anaerobic Granular Sludge Reactors. Environmental Technology, 2003. 24(5): p. 573 - 587.

186. Vallee, B.L. and D.D. Ulmer, Biochemical effects of mercury, cadmium, and lead. Annual Review of Biochemistry, 1972. 41(10): p. 91-128.

Page 46: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

72

72

187. Sterritt, R.M. and J.N. Lester, Interactions of heavy metals with bacteria. Science of the Total Environment, 1980. 14(1): p. 5-17.

188. Zayed, G. and J. Winter, Inhibition of methane production from whey by heavy metals--protective effect of sulfide. Applied Microbiology and Biotechnology, 2000. 53(6): p. 726-31.

189. Mosey, F.E., J.D. Swanwick, and D.A. Hughes, Factors affecting the availability of heavy metals to inhibit anaerobic digestion. Water Pollution Control, 1971. 70(6).

190. Lin, C.-Y. and C.-C. Chen, Effect of heavy metals on the methanogenic UASB granule. Water Research, 1999. 33(2): p. 409-416.

191. Lin, C.-Y., Effect of heavy metals on volatile fatty acid degradation in anaerobic digestion. Water Research, 1992. 26(2): p. 177-183.

192. Lin, C.-Y., Effect of heavy metals on acidogenesis in anaerobic digestion. Water Research, 1993. 27(1): p. 147-152.

193. Wong, M.H. and Y.H. Cheung, Gas production and digestion efficiency of sewage sludge containing elevated toxic metals. Bioresource Technology, 1995. 54(3): p. 261-268.

194. Ahring, B.K. and P. Westermann, Sensitivity of thermophilic methanogenic bacteria to heavy metals. Current Microbiology, 1985. 12(5): p. 273-276.

195. Mudhoo, A. and S. Kumar, Effects of heavy metals as stress factors on anaerobic digestion processes and biogas production from biomass. International Journal of Environmental Science and Technology, 2013. 10(6): p. 1383-1398.

196. Dong, J., et al., Partitioning of Heavy Metals in Municipal Solid Waste Pyrolysis, Gasification, and Incineration. Energy & Fuels, 2015. 29(11): p. 7516-7525.

197. Liao, C., C. Wu, and Y. Yan, The characteristics of inorganic elements in ashes from a 1 MW CFB biomass gasification power generation plant. Fuel Processing Technology, 2007. 88(2): p. 149-156.

198. James, A.K., et al., Ash Management Review-Applications of Biomass Bottom Ash. Energies, 2012. 5(10): p. 3856-3873.

199. Tchounwou, P.B., et al., Heavy Metals Toxicity and the Environment. EXS, 2012. 101: p. 133-164. 200. Abdoulmoumine, N., A. Kulkarni, and S. Adhikari, Effects of temperature and equivalence ratio on

pine syngas primary gases and contaminants in a bench-scale fluidized bed gasifier. Industrial and Engineering Chemistry Research, 2014. 53(14): p. 5767-5777.

201. Rabou, L.P.L.M., et al., Tar in biomass producer gas, the Energy research Centre of the Netherlands (ECN) experience: an enduring challenge. Energy & Fuels, 2009. 23(12): p. 6189-6198.

202. Lopes, E.J., et al., Evaluating the emissions from the gasification processing of municipal solid waste followed by combustion. Waste Management, 2018. 73: p. 504-510.

203. Edwards, E.A. and D. Grbić-Galić, Anaerobic degradation of toluene and o-xylene by a methanogenic consortium. Applied and Environmental Microbiology, 1994. 60(1): p. 313-322.

204. Foght, J., Anaerobic Biodegradation of Aromatic Hydrocarbons: Pathways and Prospects. Journal of Molecular Microbiology and Biotechnology, 2008. 15(2-3): p. 93-120.

205. Marozava, S., et al., Anaerobic degradation of 1-methylnaphthalene by a member of the Thermoanaerobacteraceae contained in an iron-reducing enrichment culture. Biodegradation, 2018. 29(1): p. 23-39.

206. Cateni, B.G., Effects of feed composition and gasification parameters on product gas from a pilot scale fluidized bed gasifier. 2007, Oklahoma State University.

207. Akyol, Ç., et al., Individual and combined inhibitory effects of methanol and toluene on acetyl-CoA synthetase expression level of acetoclastic methanogen, Methanosaeta concilii. International Biodeterioration & Biodegradation, 2015. 105: p. 233-238.

73

208. Ince, O., et al., Toluene inhibition on an anaerobic reactor sludge in terms of potential activity and composition of acetoclastic methanogens. Journal of Environmental Science and Health - Part A Toxic/Hazardous Substances and Environmental Engineering, 2009. 44(14): p. 1551-1556.

209. Weelink, S.A.B., M.H.A. van Eekert, and A.J.M. Stams, Degradation of BTEX by anaerobic bacteria: physiology and application. Reviews in Environmental Science and Bio/Technology, 2010. 9(4): p. 359-385.

210. Wikandari, R., et al., Performance of semi-continuous membrane bioreactor in biogas production from toxic feedstock containing D-Limonene. Bioresource Technology, 2014. 170: p. 350-5.

211. Doran, M., Pauline, Mass transfer, in Bioprocess Engineering Principles 2012, Elsevier. p. 379-444.

212. Chandolias, K., E. Pekgenc, and M.J. Taherzadeh, Floating Membrane Bioreactors with High Gas Hold-Up for Syngas-to-Biomethane Conversion. Energies, 2019. 12(6): p. 1046.

Page 47: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

74

74

Paper I

Rapid bio-methanation of syngas in a reverse membrane bioreactor: Membrane encased

microorganisms

Page 48: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

Rapid bio-methanation of syngas in a reverse membrane bioreactor:Membrane encased microorganisms

Supansa Youngsukkasem ⇑, Konstantinos Chandolias, Mohammad J. TaherzadehSwedish Centre for Resource Recovery, University of Borås, 50190 Borås, Sweden

h i g h l i g h t s

� A reverse membrane bioreactor (RMBR) was applied for syngas bio-methanation.� The cells encased in PVDF membrane could convert syngas and produce CH4 in 1 day.� Encased cells in the RMBR performed better at 55 �C compared to 35 �C.� Addition of organic waste with syngas improved the methane productivity.

a r t i c l e i n f o

Article history:Received 3 June 2014Received in revised form 18 July 2014Accepted 19 July 2014Available online 1 August 2014

Keywords:Syngas fermentationCo-digestionMethaneCell retentionMembrane bioreactor

a b s t r a c t

The performance of a novel reverse membrane bioreactor (RMBR) with encased microorganisms for syn-gas bio-methanation as well as a co-digestion process of syngas and organic substances was examined.The sachets were placed in the reactors and examined in repeated batch mode. Different temperaturesand short retention time were studied. The digesting sludge encased in the PVDF membranes was ableto convert syngas into methane at a retention time of 1 day and displayed a similar performance asthe free cells in batch fermentation. The co-digestion of syngas and organic substances by the RMBR(the encased cells) showed a good performance without any observed negative effects. At thermophilicconditions, there was a higher conversion of pure syngas and co-digestion using the encased cells com-pared to at mesophilic conditions.

� 2014 Elsevier Ltd. All rights reserved.

1. Introduction

The global energy demand has constantly increased for severaldecades, and it has triggered the need for research and develop-ment on renewable energy sources. The potential to create renew-able energy from waste material, including municipal solid waste(MSW), industrial waste, agricultural waste, andwaste by-productshas been developed (Fatih et al., 2011). Biogas or bio-methane is arenewable energy source with several applications in e.g., car fuel,heating, cooking, or electricity production. Biogas mainly consistsof methane and carbon dioxide but may also contain minor impu-rities of other components (Deublein and Steinhauser, 2008).

There are different types of wastes used for methane production,which can be classified as easily degradablewastes, hard degradablewastes, and non-degradable wastes. In general, to obtain methanefrom easily degradable wastes such as food wastes, a biochemical

approach called an anaerobic digestion process has been employed.On the other hand, the recalcitrance of the hardly degradable mate-rials, such as crystalline cellulose and non-degradable materialssuch as lignin and plastic wastes cannot be decomposed by themicroorganisms in an anaerobic digestion process (Nizami et al.,2009). Thermochemical processes however have the potential toconvert this kind of wastes as well as non-degradable residues fromthe digestion process intomethane. In this approach, feedstocks canbe thermally gasified into intermediate gases, called syngas,through a partial oxidation process, at a relatively high temperature.Syngas or synthesis gas primarily contains carbon monoxide (CO),hydrogen (H2), and carbon dioxide (CO2). Raw syngas can be synthe-sized intomethane by the use of metal catalysts, first introduced byFranz Fischer and Hans Tropsch. However, the high manufacturingcost and challenges of the toxic impurities for the catalysts limitits economic feasibility (Fatih, 2009; Fatih et al., 2011).

Anaerobic microorganisms, primarily acetogens, carboxytrophs,and methanogens are able to use CO and/or CO2 sources as carbonsources and use H2 as an energy source for their metabolism andproduce different bio-products (Daniell et al., 2012). Thus, the

http://dx.doi.org/10.1016/j.biortech.2014.07.0710960-8524/� 2014 Elsevier Ltd. All rights reserved.

⇑ Corresponding author. Tel.: +46 33 435 4608; fax: +46 33 435 4008.E-mail address: [email protected] (S. Youngsukkasem).

Bioresource Technology 178 (2015) 334–340

Contents lists available at ScienceDirect

Bioresource Technology

journal homepage: www.elsevier .com/locate /bior tech

Page 49: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

combination of thermal and biological processes for the conversionof feedstocks into biofuel has been explored. This alternativemethod can generate a variety of products including methane froma wide variety of materials (Daniels et al., 1977; Kimmel et al.,1991; Klasson et al., 1990, 1992). Using the fermentation processesand microorganisms offers several benefits over the catalytic pro-cess, such as being a more specific process, resulting in higheryields, having a lower energy consumption, being environmentallyfriendly, and having better robustness (Mohammadi et al., 2011;Munasinghe and Khanal, 2010).

The bio-methanation of CO, H2, and CO2 in the syngas by themicroorganisms can be performed in two pathways (Kimmelet al., 1991). The first one utilizes an acetate pathway as a methaneprecursor. Thereafter, methanogenic microorganisms convert ace-tate into methane. The other pathway utilizes the H2/CO2 pathway.CO can be converted into CO2 by the microorganisms (Henstraet al., 2007). The H2 and CO2 produced and initially present inthe syngas are converted into methane by some methanogenicmicroorganisms (Kimmel et al., 1991). However, this microbialflora including methanogens requires a long retention time, sincetheir growth rate is very low. Methanogenic microorganisms arealso very sensitive to the process conditions; hence, the cells areeasily washed out from the digester at high dilution rates. For thesereasons, the population size of the microbial cells is easily reducedresulting in a decreased methane production. Moreover, fermenta-tion processes with a low cell density need long start-up periods,and larger digesters are required for proper function, which meansthat the capital cost is high. To achieve a high process efficiency inconverting syngas into methane using microorganisms in a fer-mentation process, retaining the microbial cells inside a compactreactor might be a solution to overcome these problems (Klassonet al., 1992). Immobilized cell technology, meaning the confine-ment of cells in a specific region or matrix, has been widely usedin a variety of laboratory experiments and industrial applications(Raymond et al., 2004). These systems have been used to solveproblems encountered in conventional bioreactors using sus-pended cell culture, including low biomass concentrations, lowbiomass productivity and product formation, inefficiency in con-tinuous production, low stability to sudden fluctuation, and limiteddilution rate in the case of continuous operation (Chinnayelka andMcShane, 2004).

A membrane bioreactor (MBR) is the combination of a mem-brane process such as microfiltration with a suspended growthbioreactor (Judd and Judd, 2011). MBRs have been widely usedfor cell retention, especially in the municipal and industrial waste-water treatment. For example, MBR was studied by Badani et al.(2005) to treat textile wastewater, and it was found to performperfectly compared to the conventional system. Kanai et al.(2010) employed a submerged anaerobic membrane bioreactorprocess in the anaerobic digestion of food wastes. They found thatMBR enhanced the degradation of wastes, the rejection of toxics,and the reactor volumes could be scaled down compared to theconventional system. This technology has several benefits overthe conventional systems such as having low energy consumption,being environmentally friendly, and being easily implemented.Using MBR to retain the microbial cells for the enhancement ofthe bio-methanation of syngas could be an interesting system.However, currently available commercial MBR processes employedin biological process are designed to filter the liquid through themembranes, retaining the cells in the reactor to obtain a clarifiedand disinfected product. It is this type of MBR (the biomass rejec-tion MBR), which is in primary focus of the most research in thisfield (Judd and Judd, 2011). In order to increase the conversionefficacy and bio-productivity, a new technique for retaining thecells inside the bioreactor is probably necessary for efficient bio-methanation systems of syngas.

The current work focuses on applying the reverse membranebioreactor (RMBR) to retain the cells for the bio-conversion processof synthesis gas into methane. RMBR used in this work, the liquidpermeate is not actively passes through the membrane, but thesubstrate diffuse through the membrane and the metabolic prod-ucts diffuse back to the medium. It is a novel technique that usesmembranes (PVDF) to enclose the microbial cells completelyfollowed by immersion into a bioreactor in order to retain a highcell density in the bioreactor and increase the methane productiv-ity, without the risk of a clogged outlet. In this work, the efficiencyof the RMBR contained the encased cells and the resistance of usingthe membrane as a barrier for the syngas fermentation was testedcompared to the free cells, in repeated batch mode with a short-ened retention time and different temperatures. Moreover, thepossibility of enhancing the methane productivity with additionof organic substances as a co-substrate with syngas using theRMBR was also examined.

2. Methods

2.1. Anaerobic culture, medium, and syngas

An anaerobic culture was obtained from a 3000 m3 municipalsolid waste digester, operating under thermophilic (55 �C) condi-tions (Borås Energy and Environment AB, Sweden). The inoculumwas incubated at 55 �C for 3 days to keep the bacteria active whileconsuming the carbon source provided by the inoculum. Afterincubation, the inoculum was filtered through a sieve (1 mm poresize) to remove the large particles. The digesting sludge was thencentrifuged at 31,000�g for 15 min to separate the solid inoculum(85% TS), the methane-producing microorganisms, which wereloaded into the membrane sachets. Two kinds of synthetic mediumsolutions for two different experiments were prepared. One wasfor the co-digestion; the synthetic organic medium contains ace-tate, propionate, butyrate, and vitamins solution (basal medium)with a ratio of 3:1:1:1 (Osuna et al., 2003). The other was thevitamin solution, that is, the basal medium for the pure syngas fer-mentation. Medium solutions were buffered to pH 7.0 ± 0.2 withNaHCO3 (Isci and Demirer, 2007). Synthetic syngas containing CO(55% mole), H2 (20% mole), and CO2 (10% mole) (Kimmel et al.,1991) was purchased from AGA gas (Borås, Sweden).

2.2. Membrane sachet preparations and cell containment procedure

The cell entrapment in the membranes was performed follow-ing a previously described method (Youngsukkasem et al., 2012,2013). Flat plain PVDF (polyvinylidene fluoride, Durapore�) mem-branes (Thermo Fisher Scientific Inc., Sweden) were used as syn-thetic membranes. The PVDF membrane filters were hydrophilewith the pore size, thickness and diameter of 0.1 lm, 125 lmand 90 mm, respectively. The membranes were cut into rectangu-lar shapes of 6 � 6 cm and folded to create membrane pockets of3 � 6 cm2. They were then heat-sealed (HPL 450 AS, Hawo,Germany) on two sides with heating and cooling times of 4.5 and4.5 s, leaving one side open for the insertion of the inoculum. Solidsludge inoculum (3 g per sachet) was then injected carefully intothe synthetic membrane sachets, and the fourth side sealed. Thesachets containing the inoculum were used immediately for thebio-methanation of the syngas and the co-digestion process.

2.3. Repeated batch fermentation process

The repeated batch fermentation processes (with cell reuse andmedium replacement) were performed in order to preliminarilyexamine the performance of the RMBR (microorganisms encased

S. Youngsukkasem et al. / Bioresource Technology 178 (2015) 334–340 335

in the synthetic membrane) for the syngas bio-methanation andthe co-digestion of syngas and organic medium at different reten-tion times. The schematic diagram of an experiment shows inFig. 1. In the experiment with the syngas bio-methanation, eachreactor contained one sachet of inoculum and 30 mL of basalmedium. For the experiment with the co-digestion process, eachdigester contained one sachet of inoculum and 30 mL of syntheticorganic medium. The reactors used were serum glass bottles with118 mL working volume, closed with butyl rubber seals and alumi-num caps. The headspace of each bottle was flushed with syngas toobtain a sufficient amount of gas substrate at 1 atm. Reactors wereinclined and carried out in a water bath with constant agitation of100 rpm in order to provide good gas–liquid mass transfer. Anaer-obic fermentations at temperatures of 35 �C and 55 �C were exam-ined. The retention time was gradually shortened: 4, 4, 2, 2, and1 day, respectively. The encased cells were reused; the syngaswas exchanged with fresh syngas at 1 atm, and necessary nutrientswere replaced prior to the start of every new batch. Anaerobicfermentations of the free cells with syngas were performed for9 days of retention time in parallel in order to examine theefficiency of the encased cells for the bio-methanation of syngas,especially, the membrane resistance, under identical conditionsas the reference.

2.4. Analytical methods

Methane, hydrogen, carbon monoxide, and carbon dioxide weremeasured regularly, using a gas chromatograph (Perkin-Elmer,U.S.A.), equipped with a packed column (CarboxenTM 1000, SUPE-LCO, 60 � 1.800 OD, 60/80 Mesh, U.S.A.) and a thermal conductivitydetector (Perkin-Elmer, U.S.A.) with an injection temperature of200 �C. The carrier gas was nitrogen, with a flow rate of30 mL/min at 75 �C. A 250 lL gastight syringe (VICI, PrecisionSampling Inc., U.S.A.) was used for the gas sampling. The obtainedpeak area was compared with a standard gas analyzed at the samecondition (STP, 273.15 K and 101.325 kPa). The volatile fatty acids(VFA) were analyzed using a gas chromatograph (Auto System,Perkin-Elmer, U.S.A.) equipped with a capillary column (ZebronZB-WAX plus, Polyethylene glycol (PEG), 30 m � 0.25 mm �0.25 lm, U.S.A.) and a flame ionized detector (Perkin-Elmer,U.S.A.) with an injection and detection temperature of 250 �C and300 �C, respectively. The carrier gas was nitrogen, with a flow rateof 2 mL/min at a pressure of 20 psi. The experiment was performed

in triplicate and the results were presented as mean ± standarddeviation.

3. Results and discussion

3.1. The performance of the RMBR (digesting microorganisms encasedin the PVDF membranes) compared to the free cells for the bio-methanation of syngas

During the anaerobic bio-methanation process of syngas, it wasobserved that the encased microorganisms performed smoothly.Sachets containing the cells stayed intact and no leakage of cellsoccurred throughout the anaerobic conversion process.

Fig. 2 shows the performance of the encased cells compared tothe free cells in the anaerobic bio-methanation of syngas. Theencased microbial cells had a similar accumulated methane levelcompared to the free cells in the thermophilic fermentation inthe 9-day batch. Methane was rapidly accumulated in the reactorswith both the encased and the free cells at thermophilic conditions(55 �C), from the first day until the 3rd day of fermentation. There-after, the trend of methane production was stable until the last daydue to slow production (Fig. 2a). The accumulated methane pro-duced on the 9th day for the encased cells and the free cells at55 �C and 35 �C were 0.72, 0.41, 0.51, and 0 mmol, respectively.The amount of H2 and CO decreased dramatically during the first3 days. Under the thermophilic conditions, the free cells showedbetter performance in rapid assimilation of the syngas comparedto the encased cells. CO and H2 were completely consumed bythe free cells already in the 2nd day of bio-methanation. The anaer-obic digesting sludge used as the inoculum contained differentspecies of microorganisms, and used for biogas production, as anactive source of methane-producing microorganisms. In addition,it has been found to be a potential source of microorganisms forthe syngas fermentation in different bioprocess systems (Sipmaet al., 2003). The individual microbial cells (free cells) most likelyhave better contact to the medium with dissolved gases in the fer-mentation system compared to the compact cells encased in themembranes. However, cells encased in the PVDF membranes con-verted all H2 into products in 4 days while CO required a longertime to be completely used up by the encased cells. The trend ofCO2 consumption was stable for all treatments (Fig. 2d).

This experiment revealed that microorganisms encased in thePVDF membrane displayed a similar performance compared tothe free cells in syngas bio-methanation, but the conversion rateof syngas was found to be a bit slower than for the free cells with

A reverse membrane bioreactor (RMB)

(Encased cells)Free microbial cells

Pure syngas fermenta�on

Co-digestion of syngas and organic

substances

Conditions:Shortened retention times and different

temperatures

Analyzation

CO, H2, CO2, CH4and VFAs

Fig. 1. The schematic diagram of an experiment.

Fig. 2. Anaerobic bio-methanation of syngas using the RMBR (encased cells)compared to free cells. (a) Methane, (b) Hydrogen, (c) Carbon monoxide, and (d)Carbon dioxide. Symbols: Encased cells at 35 �C ( ), Free cells at 35 �C ( ),Encased cells at 55 �C ( ), and Free cells at 55 �C ( ).

336 S. Youngsukkasem et al. / Bioresource Technology 178 (2015) 334–340

Page 50: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

the same process conditions. However, the results indicate thatusing the PVDF membrane to retain microbial cells for bio-methanation of syngas is feasible. Hence, it would be interestingto investigate this novel technique in a long-term process.

3.2. The efficiency of the RMBR (encased digesting microorganisms) inrepeated batch process for bio-methanation of syngas

3.2.1. Effect of temperaturesFig. 3 shows the performance of the encased cells in the anaer-

obic bio-methanation of syngas at different retention times andtemperatures compared to the control treatments with no syngas,under the same conditions. The result shows that at the firstretention time of 9 days, methane was produced continuouslywhile the syngas concentration decreased. The remarkable resultshows that methane was produced immediately from the firstday of fermentation in all conditions. In addition, the interestingresult from this experiment was that the encased cells, at bothtemperature conditions, produced more methane than the controlswith no syngas at the same conditions. Furthermore, it wasobserved that under thermophilic conditions of 55 �C, the encasedcells had the highest methane production of 1.53 mmol on the 9thday compared to the mesophilic conditions at 35 �C and the con-trols at both temperatures (0.78, 0.81, and 0.37 mmol, respec-tively). Kundiyana et al. (2011) studied the syngas fermentationusing microbial cells at different temperatures and found that amesophilic temperature (35 �C) was the best condition for enhanc-ing the gas solubility in the liquid medium and consequentlyenhancing the bio-productivity. However, in the current experi-ment, under mesophilic conditions (35 �C), the encased cells pro-duced less methane than those at thermophilic conditions. Thismeans that the gas solubility may not be the main obstacle in

increasing the process efficiency of syngas fermentation. The nat-ure of the fermenting microorganisms, which might work betterat thermophilic conditions, also plays an important role for thisprocess. In this experiment, the active inoculum obtained from athermophilic biogas plant is the likely reason for the improvedbio-conversion of syngas. Thermophilic conditions generally stim-ulate the microbial metabolism compared to lower temperatures.Thus, faster bio-methanation of syngas occurred at thermophilicconditions with the community of microbial cells originating fromthe thermophilic biogas plant.

Both H2 and CO have a lower solubility in the liquid phase com-pared to CO2, but the amount of syngas, including H2 and CO(Fig. 3b and c), decreased continuously from the first day, at thesame time as methane was produced (Fig. 3a). The H2 concentra-tion of 0.7 mmol at 55 �C anaerobic condition at the first day haddecreased to 0.1 mmol at the 5th day and thereafter, was verylow during the remainder of the fermentation period (Fig. 3b). At35 �C, the encased cells had a lower conversion efficiency of H2

than at thermophilic condition. Here, the same amount of H2

(0.7 mmol) decreased slowly to 0.06 mmol on the 9th day offermentation. However, the encased cells had no problems to takeup H2 for methane formation. H2 is usually contained in syngas atdifferent concentrations depending on the composition of the bio-mass used for the thermal gasification process and also on the pro-cess parameters (Fatih, 2009). To produce methane, methanogenicmicroorganisms such as Methanothermobacter thermoautotrophicus(Sipma et al., 2003) utilize H2 together with CO2. However, precur-sors for methane-producing microorganisms, such as acetate, canalso be formed by microbial cells such as Acetoanaerobium noteraeusing H2.

With carbon monoxide (Fig. 3c), no problems regarding theinhibitory effects on the cells were observed during the fermenta-tion process. The encased cells performing at thermophilic condi-tions were able to also consume CO faster than the cells atmesophilic conditions. The CO (1.8 mmol) was completely assimi-lated on the 7th day at 55 �C. At 35 �C, the same amount of CO wascompletely used up by the encased cells by the 9th day. In order tomimic the raw syngas, CO2 was added as one of the main gases forthis experiment. The result shows that the production trends ofCO2 by the encased cells using syngas (35 and 55 �C) were quitestable (Fig. 3d). It may be because of CO2, even as a metabolicproduct (cf. controls on Fig. 2d). The amount of CO2 at the condi-tions of 35 �C and 55 �C were in the ranges of 0.25–0.50 and0.33–0.59 mmol, respectively. To formmethane, the microbial cellsprobably used CO2 from their metabolisms, so they did not needthe CO2 supplied from the syngas. However, CO2 accumulationhad no negative effect on the bio-methanation of syngas in thisexperiment.

3.2.2. Effect of short retention timesThese results reveal that anaerobic microorganisms encased in

the PVDF membrane were able to convert syngas into methaneat the retention time of 9 days. However, to investigate the effi-ciency of the encased cells in shorter periods of bio-methanation,the experiment was thereafter operated in repeated batch mode,in which the encased cells were reused and new syngas wasreplaced at the beginning of every new batch. The batch timewas gradually shortened to 4, 4, 2, 1, and 1 day, respectively. Thecompositions of the syngas and methane were analyzed regularly.

The results showed that the encased cells still performedefficiently, and no negative effect was observed during the bio-methanation of the syngas. In addition, the encased microbial cellshad better conversion efficiency of the syngas at anaerobic thermo-philic conditions than at mesophilic conditions throughout allbatches (Fig. 3). Under thermophilic anaerobic conditions, COwas completely used up by the encased cells on the 3rd day of

Fig. 3. Performance of the RMBR in anaerobic repeated batch bio-methanation ofsyngas. (a) Methane production, (b) Hydrogen concentration, (c) Carbon monoxideconcentration, and (d) Carbon dioxide concentration. Symbols: Syngas at 35 �C( ), Blank at 35 �C ( ), Syngas at 55 �C ( ), and Blank at 55 �C ( ).

S. Youngsukkasem et al. / Bioresource Technology 178 (2015) 334–340 337

the batch. When the retention time was gradually shortened to2 days, CO was still completely used up. With a retention time of1 day, the encased cells used CO, but it was not enough time forthe encased cells to completely use all CO (Fig. 3c). For H2, theamount decreased continuously from the beginning of the batches,especially, at the retention time of 4 days. Nevertheless, the reten-tion times of 4, 2, and 1 day were not enough for the encased cell toentirely convert the H2 into products. The trend for the CO2 levelwas found to be the same throughout the process, even whenthe batch time was shortened. The accumulated methane produc-tion of the encased cells in thermophilic bio-methanation withbatch times of 4, 4, 2, 1, and 1 day were 0.90, 0.88, 0.73, 0.35,and 0.35 mmol, respectively. The most interesting point from thisexperiment was that the reused encased cells performed well,and were still able to convert syngas into methane continuouslywhen the retention time was shortened. The accumulated methaneproduction was lower compared to the first period of 9 days. How-ever, a decrease was expected due to the used up methane poten-tial of the inoculum.

All these results indicate that the digesting sludge retained inthe hydrophilic PVDF membrane was able to convert syngas intomethane. Furthermore, the encased cells were easy to handle whenreused for the repeated batch fermentation process. HydrophilicDurapore� PVDF membrane allowed all gases, including syngas(substrate) and methane (product) together with the necessaryliquid nutrients for the microbial cells to diffuse through the mem-brane layer. Consequently, the microbial cells encased in the PVDFmembranes were able to take up syngas and produce methane.Thermophilic anaerobic conditions were found to enhance the con-version efficiency of syngas and facilitate faster methane produc-tion compared to at mesophilic conditions.

3.3. Enhancing the methane productivity using the encasedmicroorganisms in simultaneous bio-methanation of syngas andorganic substances

In anaerobic digestion processing of biomass, co-digestion hasbeen proven to be a powerful process due to its advantages in theimprovement of productivity and process efficiency. In this system,different kinds of biomass are put together in the same digesterand digested simultaneously in an often synergistic manner(Abouelenien et al., 2014; Long et al., 2012). CO, H2, and CO2 in syn-gas, the intermediate gas, can be used for the metabolism of themicroorganisms. However, syngas has a low energy, which maynot be sufficient for the microbial cells to produce methane, thus,to increase the efficacy of the encased cells in bio-methanation ofsyngas, a simultaneous fermentation process for the syngas andorganic substance to produce methane using the encased cellswas studied. In this experiment, the digesting sludge was encasedin the PVDF membranes and used as an inoculum. A syntheticorganic medium containing: acetate, propionate, and butyrate,and syngas, containing CO, H2 and CO2, were used as a co-substrate.Mesophilic and thermophilic anaerobic conditions were examinedand performed in repeated batch at different retention times. Syn-gas concentrations, total volatile fatty acids (VFA), and methaneproduction were analyzed regularly. Syngas bio-methanation bythe encased cells in the same condition was used for comparison.

The results show that the encased cells performed efficiently inthe co-biomethanation of syngas and organic substances. Thesachets got swollen during the anaerobic fermentation process,due to the high pressure of the methane produced. This means thatthe methanogenic microbes encased in the PVDF membranes hadobtained sufficient substrates for product formation. Here, thethermophilic anaerobic conditions were found to be advantageouscompared to the mesophilic conditions. Digesting sludge encasedin the hydrophilic PVDF membranes was able to perform in the

co-digestion process without any observed negative effects. Inthe first period of 9 days, the trends of methane production bythe encased cells in the co-digestion and in the pure syngasfermentation were similar. The highest accumulated methane pro-duction was obtained by the encased cells in the co-digestion andpure syngas fermentation (0.72 and 0.79 mmol) at thermophilicconditions (Fig. 4a). Looking at the gas amounts in the same period,H2 and CO were completely consumed in both processes by theencased cells on the 9th day. CO2 was consumed a bit at the begin-ning and was thereafter stable.

Total volatile fatty acids (VFA) in the liquid nutrients were alsomeasured on the last day of the different batches. After the firstbatch, with retention time of 9 days, the VFA levels decreased from0.97 g/L of pure syngas fermentation and 8.05 g/L of co-digestionunder thermophilic conditions at the first day to 0.08 and 0.05 g/L(Table 1). Hence, the VFA utilizations were 92% and 99%,respectively. These results reveal that the encased cells canconsume syngas simultaneously with the digestion of organicsubstrates. Nevertheless, the performance of the encased cells inthe co-digestion process did not show that the co-digestionresulted in a higher methane production during the first batchperiod (Fig. 4a). Thus, repeated anaerobic batch digestions withshorter retention times were performed.

When the encased cells were reused, the substrates (syngas andorganic solution) were exchanged with new substrates, and theretention time was shortened to 4, 4, 2, and 1 days, interestingresults were observed. It was found that the methane producedfrom the thermophilic co-digestion (55 �C) showed an increasingtrend (Fig. 4a), even as the retention time was shortened. Themicrobial cells encased in the hydrophilic PVDF membranes

Fig. 4. Comparison of syngas bio-methanation and co-digestion process of syngasand organic substances by the RMBR. (a) Methane, (b) Hydrogen, (c) Carbonmonoxide, and (d) Carbon dioxide. Symbols: Pure syngas at 35 �C ( ), Pure syngasat 55 �C ( ), Syngas + organic substances at 35 �C ( ), and Syngas + organicsubstances at 55 �C ( ).

338 S. Youngsukkasem et al. / Bioresource Technology 178 (2015) 334–340

Page 51: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

produced the highest amount of methane from 4, 4, and 2 days of0.42, 1.41, and 1.92 mmol in thermophilic co-digestion comparedto the others. The encased microbial cells themselves may alreadyhave adapted from the first batch, so that they could perform bet-ter and produce methane faster in the following batch periods.However, the methane level was lower when the retention timewas shortened to 1 day (1.14 and 1.59 mmol for the two batches,respectively) due to a too limited process time. In contrast, themethane production from the encased cells using only syngas asa substrate was lower than in the co-digestion process, and theproduced amounts were similar in all batches except for the short-est batches, where the production rate was obviously too slow.

After 9 days of fermentation, a new batch was started and theretention time was shortened, then H2 was completely consumedby the encased cells in 4 days of two batches for thermophilicco-digestion process (Fig. 4b). The encased cells were still able touse H2 for their metabolism at the shorter retention time, but H2

was not completely used up in the two and one day batches. COwas found to be consumed more rapidly by the encased cells inthe thermophilic pure syngas fermentation compared to in theco-digestion (Fig. 4c). It seems that the microorganisms encasedin the PVDF membranes preferred to use the other sources of sub-strates compared to CO. The encased microorganisms managed tocompletely convert the CO that was present in 2 days. For the CO2

values in this experiment (Fig. 4d), it was shown to have an inter-esting trend. After the first period of 9 days, the CO2 was consumedrapidly by the encased cells under both mesophilic and thermo-philic conditions of the co-digestion processes, but not in the sys-tems with only syngas fermentation. It is presumed that theencased digesting sludge probably contains specific microorgan-isms such as M. thermoautotrophicus, which are able to utilizeCO2 together with gases such as H2 to form precursors (acetate)or methane. Furthermore, the co-digestion process may providesuitable conditions for this microbial group, which was not thecase with only syngas fermentation.

According to the methane production and syngas consumption,the VFA concentration in the effluents from the co-digestion pro-cess was found to decrease (Table 1). The lower VFA concentrationsin the system with the encased cells in thermophilic co-digestioncompared to mesophilic conditions show that the encased cellshad spent more organic substances for their metabolism, and ithas been shown that the hydrophilic PVDF membrane allowedliquid organic substrates to diffuse through the membrane layerfor the metabolism of the microorganism in the digestion process(Youngsukkasem et al., 2013). However, the cells in co-digestionwere probably inhibited by the VFA in the first 2 and 3 batchesand thereafter, they adapted and could quickly utilize the VFA aswell as the syngas (Table 1).

The results presented here reveal that the microbial cellsencased in the hydrophilic PVDF membranes were capable of

simultaneous fermentation of the syngas and organic substratesfor methane production, without any detected negative effects.These successful results show the feasibility of applying this noveltechnique in combined anaerobic organic digestion with syngasfermentation, in order to increase the methane productivity. Inaddition, the PVDF membranes turned out to be an interestingsupporting material in this experiment. It was able to retain themicrobial cells in the bioreactors, and at the same time it was ableto allow both organic and gas substrates to pass through, whichcan provide the necessary nutrients for the metabolism of themicrobial cells to produce methane.

4. Conclusion

Retaining microorganisms in the reverse membrane bioreactor(RMBR) using the PVDF membranes was a successful approachfor bio-methanation of syngas, as well as simultaneous fermenta-tion of syngas and organic substances. The PVDF membranesallowed the liquid and gas diffusion through the membrane sur-face. The encased cells in RMBR could convert CO, H2, and CO2

and produce methane in 1 day. Thermophilic conditions (55 �C)enhanced the syngas fermentation, and the co-digestion usingthe encased cells improved the methane productivity. However,to develop RMBR system for the industrial scale, RMBR performsunder continuous bio-methanation of syngas should be furtherdeveloped.

Acknowledgement

This work was financially supported by Swedish ResearchCouncil.

References

Abouelenien, F., Namba, Y., Kosseva, M.R., Nishio, N., Nakashimada, Y., 2014.Enhancement of methane production from co-digestion of chicken manure withagricultural wastes. Bioresour. Technol. 159, 80–87.

Badani, Z., Ait-Amar, H., Si-Salah, A., Brik, M., Fuchs, W., 2005. Treatment of textilewaste water by membrane bioreactor and reuse. Desalination 185 (1–3), 411–417.

Chinnayelka, S., McShane, M.J., 2004. Resonance energy transfer nanobiosensorsbased on affinity binding between apo-enzyme and its substrate.Biomacromolecules 5 (5), 1657–1661.

Daniell, J., Köpke, M., Simpson, S., 2012. Commercial biomass syngas fermentation.Energies 5 (12), 5372–5417.

Daniels, L., Fuchs, G., Thauer, R.K., Zeikus, J.G., 1977. Carbon monoxide oxidation bymethanogenic bacteria. J. Bacteriol. 132 (1), 118–126.

Deublein, D., Steinhauser, A., 2008. Biogas from Waste and Renewable Resources.Wiley-VCH Verlag GmbH & Co., Germany.

Fatih, D.M., 2009. Biorefineries for biofuel upgrading: a critical review. Appl. Energ.86 (1(0)), S151–S161.

Fatih, D.M., Balat, M., Balat, H., 2011. Biowastes-to-biofuels. Energ. Convers.Manage. 52 (4), 1815–1828.

Henstra, A.M., Sipma, J., Rinzema, A., Stams, A.J.M., 2007. Microbiology of synthesisgas fermentation for biofuel production. Curr. Opin. Biotech. 18 (3), 200–206.

Isci, A., Demirer, G.N., 2007. Biogas production potential from cotton wastes. Renew.Energ. 32 (5), 750–757.

Judd, S., Judd, C., 2011. The MBR Book – Principles and Applications of MembraneBioreactors for Water and Waste Water Treatment, second ed. Elsevier Ltd.,Burlington.

Kanai, M., Ferre, V., Wakahara, S., Yamamoto, T., Moro, M., 2010. A novelcombination of methane fermentation and MBR – Kubota submergedanaerobic membrane bioreactor process. Desalination 250 (3), 964–967.

Kimmel, D.E., Klasson, K.T., Clausen, E.C., Gaddy, J.L., 1991. Performance of trickle-bed bioreactors for converting synthesis gas to methane. Appl. Biochem.Microbiol. 28–29 (1), 457–469.

Klasson, K.T., Ackerson, M.D., Clausen, E.C., Gaddy, J.L., 1992. Bioconversion ofsynthesis gas into liquid or gaseous fuels. Enzyme. Microb. Tech. 14 (8), 602–608.

Klasson, K.T., Elmore, B.B., Vega, J.L., Ackerson, M.D., Clausen, E.C., Gaddy, J.L., 1990.Biological production of liquid and gaseous fuels from synthesis gas. Appl.Biochem. Microbiol. 24–25 (1), 857–873.

Kundiyana, D.K., Wilkins, M.R., Maddipati, P., Huhnke, R.L., 2011. Effect oftemperature, pH and buffer presence on ethanol production from synthesisgas by ‘‘Clostridium ragsdalei’’. Bioresour. Technol. 102 (10), 5794–5799.

Table 1The concentrations of total volatile fatty acids (VFAs) from the effluent of each reactorat the start-up and at the end of the different batches.

Retention time (Days) Total volatile fatty acids (g/L)

Pure syngasfermentation

Co-digestionprocess

35 �C 55 �C 35 �C 55 �C

0 0.97 ± 0.00 0.97 ± 0.00 8.05 ± 0.00 8.05 ± 0.009 0.22 ± 0.06 0.08 ± 0.02 0.14 ± 0.03 0.05 ± 0.024 0.54 ± 0.14 0.21 ± 0.17 3.28 ± 0.74 3.43 ± 0.134 0.43 ± 0.05 0.19 ± 0.06 4.26 ± 0.19 2.94 ± 1.532 0.61 ± 0.26 0.15 ± 0.04 4.18 ± 1.83 1.41 ± 0.171 0.37 ± 0.01 0.39 ± 0.41 4.24 ± 0.73 3.27 ± 1.501 0.29 ± 0.04 0.28 ± 0.17 4.83 ± 0.57 3.27 ± 0.92

S. Youngsukkasem et al. / Bioresource Technology 178 (2015) 334–340 339

Long, J.H., Aziz, T.N., F.L.d.l., Reyes Iii, Ducoste, J.J., 2012. Anaerobic co-digestion offat, oil, and grease (FOG): a review of gas production and process limitations.Process. Saf. Environ. 90 (3), 231–245.

Mohammadi, M., Najafpour, G.D., Younesi, H., Lahijani, P., Uzir, M.H., Mohamed,A.R., 2011. Bioconversion of synthesis gas to second generation biofuels: areview. Renew. Sust. Energ. Rev. 15 (9), 4255–4273.

Munasinghe, P.C., Khanal, S.K., 2010. Biomass-derived syngas fermentation intobiofuels: opportunities and challenges. Bioresour. Technol. 101 (13), 5013–5022.

Nizami, A.-S., Korres, N.E., Murphy, J.D., 2009. Review of the integrated process forthe production of grass biomethane. Environ. Sci. Technol. 43 (22), 8496–8508.

Osuna, M.B., Zandvoort, M.H., Iza, J.M., Lettinga, G., Lens, P.N., 2003. Effects of traceelement addition on volatile fatty acid conversions in anaerobic granular sludgereactors. Environ. Technol. 24 (5), 573–587.

Raymond, M., Neufeld, R., Poncelet, D., 2004. Encapsulation of brewers yeast inchitosan coated carrageenan microspheres by emulsification/thermal gelation.Artif. Cells Blood Subst. Immobil. Biotechnol. 32 (2), 275–291.

Sipma, J., Lens, P.N.L., Stams, A.J.M., Lettinga, G., 2003. Carbon monoxide conversionby anaerobic bioreactor sludges. FEMS Microbiol. Ecol. 44 (2), 271–277.

Youngsukkasem, S., Akinbomi, J., Rakshit, S.K., Taherzadeh, M.J., 2013. Biogasproduction by encased bacteria in synthetic membranes: protective effects intoxic media and high loading rates. Environ. Technol., 1–8.

Youngsukkasem, S., Rakshit, K.S., Taherzadeh, M.J., 2012. Biogas production byencapsulated methane producing bacteria. Bioresources 7 (1), 56–65.

340 S. Youngsukkasem et al. / Bioresource Technology 178 (2015) 334–340

Page 52: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

75

Paper II

Syngas Biomethanation in a Semi-Continuous Reverse Membrane Bioreactor (RMBR)

Page 53: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

fermentation

Article

Syngas Biomethanation in a Semi-ContinuousReverse Membrane Bioreactor (RMBR)Supansa Y. Westman *, Konstantinos Chandolias and Mohammad J. Taherzadeh

Swedish Centre for Resource Recovery, University of Borås, Borås 50190, Sweden;[email protected] (K.C.); [email protected] (M.J.T.)* Correspondence: [email protected]; Tel.: +46-70-487-7100

Academic Editor: Thaddeus EzejiReceived: 22 January 2016; Accepted: 21 March 2016; Published: 25 March 2016

Abstract: Syngas biomethanation is a potent bio-conversion route, utilizing microorganisms toassimilate intermediate gases to produce methane. However, since methanogens have a long doublingtime, the reactor works best at a low dilution rate; otherwise, the cells can be washed out during thecontinuous fermentation process. In this study, the performance of a practical reverse membranebioreactor (RMBR) with high cell density for rapid syngas biomethanation as well as a co-substrate ofsyngas and organic substances was examined in a long-term fermentation process of 154 days andcompared with the reactors of the free cells (FCBR). The RMBR reached maximum capacities of H2, CO,and CO2 conversion of 7.0, 15.2, and 4.0 mmol/Lreactor.day, respectively, at the organic loading rate of3.40 gCOD/L.day. The highest methane production rate from the RMBR was 186.0 mL/Lreactor.dayon the 147th day, compared to the highest rate in the FCBR, 106.3 mL/Lreactor.day, on the 58th day.The RMBR had the ability to maintain a high methanation capacity by retaining the microbial cells,which were at a high risk for cell wash out. Consequently, the system was able to convert moresyngas simultaneously with the organic compounds into methane compared to the FCBR.

Keywords: syngas fermentation; methane production; semi-continuous process; reverse membranebioreactor; co-substrate; cell retention

1. Introduction

Biogas is a renewable energy source with several applications in heating, cooking, or electricityproduction. Biogas mainly consists of methane and carbon dioxide but may also contain minorimpurities of other components [1]. It can also be upgraded to about 97% biomethane, which is usedas a popular renewable form of energy such as car fuel.

Biogas or biomethane can be produced from waste resources that can be classified as easilydegradable, difficult to degrade, and non-degradable wastes. To obtain biogas from easily degradablewastes such as food wastes, a biochemical approach called anaerobic digestion process is traditionallyemployed. In this process, complex polymers of organic substances are degraded easily bymicroorganisms through the different steps of the anaerobic fermentation process [2]. In the sameway, difficult to degrade materials such as lignocelluloses from crop residues and agricultural residuescan also be converted into methane by the performance of microorganisms in the anaerobic digestionprocess. However, the recalcitrance of these difficult to degrade biomass sources, such as the coverageof lignin or the crystallinity of cellulose, limits the access of the microorganisms and enzymes forefficient digestion [3]. The hydrolysis rate can be accelerated by pretreating the substrate with chemical,physical, or biological processes before feeding the substrate to the digester [4]. Nevertheless, thepre-treatment process has been found to be an expensive and rate limiting factor [5]. Thus, thebiochemical process may face challenges in converting difficult to degrade wastes into methane.Non-degradable materials such as lignin or plastic wastes cannot be decomposed by microorganisms

Fermentation 2016, 2, 8; doi:10.3390/fermentation2020008 www.mdpi.com/journal/fermentation

Page 54: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

Fermentation 2016, 2, 8 2 of 12

in an anaerobic digestion process. Consequently, it is common to incinerate this category of wasteto energy.

On the other hand, it is possible to convert non-degradable wastes and lignocelluloses to biofuelsvia gasification to syngas and the fermentation routes. The biomass is first gasified into a gas mixturecalled syngas or synthesis gas. Syngas contains primarily CO, H2, and CO2. The syngas can beupgraded by methanization, either by using chemical catalysts (e.g., Fischer-Tropsch Synthesis) orthrough fermentation with microbial catalysts. The fermentation process offers several benefits over thechemical catalytic process, such as a more specific process, higher yields, lower energy consumption,being environmentally friendly, and having better robustness [6,7]. Syngas fermentation for methaneproduction, i.e., syngas biomethanation, is a relatively new technology and a good alternative to thefirst-generation biofuel.

The biomethanation of the syngas by microorganisms can be performed in two pathways [8]. Thefirst one utilizes an acetate pathway as a methane precursor (Reactions 1 and 2). Acetobacterium woodiiand Enbacterium limosum, for example, are microbial cells that perform this reaction. Thereafter,methanogenic microorganisms such as Methanosarcina barkeri convert acetate into methane (Reaction 3):

2CO2 ` 4H2 Ñ CH3COOH ` 2H2O Reaction 1

4CO2 ` H2O Ñ CH3COOH ` 2CO2 Reaction 2

CH3COOH Ñ CH4 ` CO2 Reaction 3

The other pathway utilizes the H2/CO2 pathway. CO can be converted into CO2

(Reaction 4) by microorganisms, for example, Methanothermobacter thermoautotrophicus andClostridium thermoaceticum [9]. The H2 and CO2 produced by reaction 4 and initially present inthe syngas are converted into methane by some microorganisms such as Methanosarcina formicicum(Reaction 5) [8]:

CO ` H2O Ñ CO2 ` H2 Reaction 4

4H2 ` CO2 Ñ CH4 ` H2O Reaction 5

However, these microbial groups, especially, methane-producing microorganisms, require a longretention time in the digester, since their growth rate is very low. In addition, they are very sensitive toharsh conditions; hence, the cells can easily be washed out from the digester at high dilution rates ina continuous process. For these reasons, the population size of the microbial cells is easily reduced,resulting in a decreased methane production and low process efficiency. Moreover, fermentationprocesses with low cell density need long start-up periods, and larger digesters are required for properfunction, which means that the capital cost is high. Retaining the microbial cells inside a compactbioreactor with a low chance of bio-fouling and abraded microorganisms might be a solution toovercome these problems.

Different bioreactor designs have been developed in order to enhance the process efficiency ofthe syngas fermentation [10–15]. For instance, continuous stirred tank reactors (CSTR) have beenfrequently employed for syngas fermentation. A high speed of agitation in the CSTR can be used toenhance the mass transfer efficiency. However, it has been observed to result in high shear stress onthe cells and high-energy consumption [16,17].

In another study, a bioreactor for syngas fermentation was connected to a hollow fiber membranemodule [16]. The system enhanced the gas-liquid mass transfer as well as allowed the microbial cellsto be attached on the membrane layer [16]. However, a long start-up period for cell attachment to theouter membrane layer was crucial. The same authors also developed a novel bioreactor configuration,called a monolithic biofilm reactor, for the same purpose. The results revealed that the performanceof the syngas fermentation process depended not only on the mass transfer efficiency, but the lowefficiency was also related to bio-fouling and abrading of the microbial cells attached to the channelof the monolithic wall [18]. Liu et al. [19] used the membrane module for the cell retention system in

Fermentation 2016, 2, 8 3 of 12

continuous syngas fermentation for ethanol production. However, cell wash out still occurred, whichwas one of the reasons for low product formation.

From these results and the nature of methane-producing microorganisms, retaining a highmicrobial cell density in the bioreactor and preventing cell washout during the syngas continuousfermentation processes are important aspects to keep the process stable and increase the methaneproductivity [15].

The reverse membrane bioreactor (RMBR) is a configuration of a bioreactor for methaneproduction. The RMBR system can be defined as a process that employs a fixed microbial growthsystem with the use of membranes. This bioreactor system was designed to enclose the cells completelyin the permeable membrane filter, while substrates in the surrounding liquid are able to diffuse throughthe membrane to the cells. The fermentation products can thereafter diffuse out to the surroundingsolution. Additional equipment for the cell recycle system is not necessary for this kind of process.This technique provided a high cell density; consequently, the process efficiency and productivity ofthe anaerobic digestion process for methane production were improved [20–23].

The current work focuses on applying the RMBR for syngas biomethanation under anaerobicsemi-continuous mode in order to retain a high cell loading and consequently, improve the processperformance. The simultaneous fermentation of syngas and organic compounds was also investigated.The performance of the system was evaluated under thermophilic anaerobic conditions and comparedto the conventional system of free cells bioreactor (FCBR) under the same conditions.

2. Materials and Methods

2.1. Anaerobic Culture, Medium, and Raw Syngas

An anaerobic culture was obtained from a 3000-m3 municipal solid waste digester, operatingunder thermophilic (55 ˝C) conditions (Borås Energy and Environment AB, Borås, Sweden). Theinoculum was incubated at 55 ˝C for 3 days to keep the bacteria active while consuming the carbonsource provided by the inoculum. After incubation, the inoculum was filtered through a sieve(1-mm pore size) to remove the large particles. The digesting sludge was then centrifuged at 31,000ˆ gfor 15 min to separate the solid inoculum (85 % Total solid (TS)) and the microorganisms, whichwere loaded into membrane sachets (Section 2.2). The synthetic medium solution contained acetate,propionate, butyrate with COD strength of 6.72 gCOD/L, and a vitamin solution (basal medium)with COD strength of 3.34 gCOD/L at a ratio of 3:1:1:1 [24]. The medium solution was buffered topH 7.0 ˘ 0.2 with NaHCO3 [25]. The artificial syngas containing CO (55% mol), H2 (20% mol), andCO2 (10% mol) [11] was provided by AGA gas AB (Borås, Sweden).

2.2. Cell Containment Procedure for the RMBR Configuration

The microbial cell encasement was performed following a previously described method [23].Flat plain hydrophilic PVDF (polyvinylidene fluoride, Durapore®) membranes (Merch Millipore Ltd.,Cork, Ireland) with a pore size, thickness, and diameter of 0.1 μm, 125 μm, and 90 mm, respectively,were used as supporting materials. The main physicochemical characteristics of the membranes are: airflow rate of 0.15 L/min.cm2, 0.5% gravimetric extractables, 70% porosity, and water flow rate greaterthan 0.33 mL/min.cm2. The membranes were cut into rectangular shapes of 6 ˆ 6 cm2 and folded tocreate membrane pockets of 3 ˆ 6 cm2. They were then heat-sealed (HPL 450 AS, Hawo, Obrigheim,Germany) on three sides with heating and cooling times of 4.5 and 4.5 s, leaving one side open for theinsertion of the inoculum. Solid sludge inoculum (3 g per sachet) was then injected carefully into thesynthetic membrane pockets, and the fourth side was subsequently sealed. The sachets containing theinoculum were used immediately for biomethanation of syngas.

Page 55: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

Fermentation 2016, 2, 8 4 of 12

2.3. Reactor Set up for Semi-Continuous Operation

The schematic illustration of the RMBR system is shown in Figure 1. The reactors used wereserum glass bottles with 600-mL working volume, closed with rubber seals and an outlet for the biogas.The RMBR contained 15 sachets of encased inoculum (a total of 45 g) and 475 mL of synthetic medium.In parallel, the same amount of solid inoculum and medium was used in a reference reactor of freecells (FCBR) and performed under the same process conditions. The thermophilic temperature wasmaintained at 55 ˘ 1 ˝C throughout the process by placing the reactors in a water bath. The syngasamount was gradually increased by flushing fresh syngas at different gas pressures. Simultaneously,the OLR was gradually increased by increasing the concentrations of the medium fed (Table 1). Duringthe fermentation process, the syngas was continuously bubbled through the reactors, from the bottomrising up to the top of the reactors, from where it was circulated to the bottom at different flow rates.Each co-substrate loading rate and gas circulation rate was maintained for different experiments(different periods of time), giving a total digestion process of 154 days.

Figure 1. Schematic illustration of the reverse membrane bioreactor (RMBR) in the semi-continuousbiomethanation process of syngas and organic substances. (a) Digesting sludge encased in PVDFmembrane; (b) Organic and nutrient inlet; (c) Syngas inlet; (d) Peristaltic pump; (e) Product outlet;(f) Warm water bath; (g) Effluent outlet; (h) Gas sampling; and (i) Data analysis.

Table 1. Parameter profiles during the simultaneous semi-continuous biomethanation process.

Periods (Days) Gas CirculationRate (mL/min)

Organic Loading(gCOD/Lreactor.day)

Syngas Amounts (mmol/Lreactor.day)

H2 CO CO2

I (0–16) 100 0.02 1.1 3.1 1.6II (17–36) 200 0.02 2.2 6.1 1.7III (37–58) 200 0.02 2.8 7.7 1.6IV (59–87) 300 0.43 1.5 4.0 0.7V (88–126) 300 0.86 1.7 4.7 1.0

VI (127–147) 300 1.70 3.8 9.4 1.9VII (148–154) 300 3.40 7.0 15.2 4.0

The parameter profiles are shown in Table 1. At the beginning of the digestion (periods I,II, and III), the syngas amount added was gradually increased. The OLR was kept relatively low,according to a very low organic content in the basal medium that was used as the vitamins and

Fermentation 2016, 2, 8 5 of 12

minerals solution during these periods. This was done in order to allow the anaerobic culture to adaptto the environment with a low risk of cell washout and to investigate the ability of both systems toresult in pure syngas fermentation. Then, from period IV to VII, the organic compounds containingacetate, propionate, and butyrate were added simultaneously with syngas. The OLR was doubled inorder to investigate the efficiency of the systems under the harsh conditions at a shorter retention time.

For the adjustment of the syngas compounds inside the reactor, syngas was placed in a 1-L gassampling bag and then the gas was fed to the reactor by a peristaltic pump. Inside the sampling bag,there was a known syngas volume that could be adjusted, and the content of each compound wasanalysed. Then, the gas was placed inside the reactor (with a known headspace), and the percentageof each syngas component was measured again. The ideal gas law was used for calculating themole number:

PV = nRT, where:P: 1 atm, adjusted with a needle after the syngas feeding; V: the headspace of the reactor (stable);

R: the ideal gas law constant; T: the temperature of the reactor; n: the unknown mole value thatwas calculated.

2.4. Analytical Methods

The composition of the gas samples (methane, hydrogen, carbon monoxide, and carbon dioxide)was measured regularly, using a gas chromatograph (Perkin-Elmer, Norwalk, CT, USA), equippedwith a packed column (CarboxenTM 1000, SUPELCO, 6’ ˆ 1.8” OD, 60/80 Mesh, Shelton, CT, USA)and a thermal conductivity detector (Perkin-Elmer) with an injection temperature of 200 ˝C. Thecarrier gas was nitrogen, with a flow rate of 30 mL/min at 75 ˝C. A 250-μL gas tight syringe (VICI,Precision Sampling Inc., Baton Rouge, LA, USA) was used for the gas sampling. The obtained peakarea was compared with a standard gas, analyzed under the same conditions (STP, 273.15˝K and101.325 kPa). The volatile fatty acids (VFAs) were analyzed using a gas chromatograph (Auto System,Perkin-Elmer) equipped with a capillary column (Zebron ZB-WAXplus, Polyethylen glycol (PEG),30 m ˆ 0.25 mm ˆ 0.25 μm, Shelton, CT, USA) and a flame ionized detector (Perkin-Elmer) with aninjection and detection temperature of 250 and 300 ˝C, respectively. The carrier gas was nitrogen, witha flow rate of 2 mL/min at a pressure of 20 psi. The experiment was performed in duplicate, and theresults were presented as the average ˘ standard deviation.

3. Results and Discussion

Syngas biomethanation is a potent bio-conversion route, utilizing microorganisms to assimilateintermediate gases to produce methane. However, the cell density in continuous bioreactor operationsplays an important role in the process efficiency. In this work, a novel application of the RMBR wasinvestigated: rapid syngas biomethanation in a semi-continuous operation at high cell loading. Theperformance of the RMBR was compared to the performance of a reactor with free cells (FCBR) underotherwise identical conditions.

During the semi-continuous biomethanation process of 154 days, the performance of the RMBRand the FCBR was investigated under seven experimental conditions (seven periods). The conditionsof each period are presented in Table 1.

3.1. Methane Production

Figure 2 and Table 2 show the trend of the methane production throughout the anaerobic process.Methane was produced from mainly pure syngas at the beginning of the process (Periods I–III),to thereafter be produced also from the co-substrates until the last day. The methane formed fromthese substrates was likely made by the acetic cleavage of acetate and the reduction of carbon dioxide.The microorganisms assimilating these substrates are mainly hydrogenotrophic methanogens andacetotrophic methanogens. It was observed that the methane production increased with the increasingconcentration of gaseous and organic substrates fed, until the maximum conversion capacity was

Page 56: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

Fermentation 2016, 2, 8 6 of 12

reached and the production decreased. This shows that the metabolism of the microorganisms occurredas expected.

Figure 2. Comparison of the methane production in the reverse membrane bioreactor (RMBR)system and the free cell bioreactor (FCBR) during the semi-continuous biomethanation of syngasand organic substances.

Table 2. Investigated parameters during the semi-continuous biomethanation process ofthe co-substrates.

PeriodsMethane Production (mL/Lreactor.day) Total VFA (g/L)

RMBR * FCBR ** RMBR * FCBR **

I 39.1 ˘ 0.4 31.4 ˘ 0.5 0.4 ˘ 0.0 0.5 ˘ 0.0II 38.0 ˘ 0.7 46.4 ˘ 0.5 0.4 ˘ 0.0 0.5 ˘ 0.0III 108.9 ˘ 0.2 106.3 ˘ 0.2 0.4 ˘ 0.0 0.5 ˘ 0.0IV 92.5 ˘ 1.5 84.4 ˘ 1.2 0.8 ˘ 0.3 1.0 ˘ 0.3V 160.8 ˘ 1.1 74.5 ˘ 1.7 0.4 ˘ 0.3 1.2 ˘ 0.7VI 186.0 ˘ 1.3 46.0 ˘ 0.6 4.4 ˘ 3.5 4.0 ˘ 3.0VII 176.5 ˘ 0.3 36.2 ˘ 0.1 7.8 ˘ 0.8 9.4 ˘ 0.5

* The reverse membrane bioreactor. ** The free cells bioreactor.

In the beginning of the experiment, a total pure syngas amount of 5.8 mmol/L.day was added tothe reactors. The amount was thereafter gradually increased throughout the operation of the reactors(Table 1). The basal medium with an organic content was fed at a low rate of 0.02 gCOD/L.dayduring the periods I, II, and III, and a similar methane production was observed in both reactors.The methane amount produced in the RMBR and the FCBR during these periods ranged between38.0–108.9 mL/L.day and 31.4–106.3 mL/L.day, respectively. These results show that under a low riskof cell wash-out condition, the encased microorganisms in the RMBR displayed a similar methaneproduction performance from syngas as the free microorganisms in the FCBR. The results also indicatethat using PVDF membranes as a cell supporting material in the current reactor setup was not a majorissue for the continuous syngas fermentation when it came to diffusion limitations. In addition, due tothe practical design of the RMBR, the problematic issues of cell clogging and cell abrading are likelyless severe.

During the periods IV, V, and VI, the organic compounds were introduced to both the reactorswhile the syngas amount added to the systems was decreased slightly in order to maintain a mildcondition for microbial adaptation. From these experiments, the microorganisms retained in theRMBR could assimilate the co-substrate perfectly with maintained efficient performance, resulting in acorresponding increase of the methane production (Table 2 and Figure 2). The maximum methaneproductivity of the RMBR during these periods was 92.5, 160.8, and 186.0 mL/L.day, respectively.On the other hand, the wash-out of the microbial cells from the FCBR was clearly occurring during

Fermentation 2016, 2, 8 7 of 12

these periods, resulting in a substantially lower methane production from the FCBR than the RMBRduring the periods V and VI (74.5 and 46.0 mL/L.day). In the RMBR, the microbial cells were stillenclosed completely in the membrane sachets, and cell abrading was not visually occurring.

A better cell retention performance of the RMBR for syngas biomethanation in a semi-continuousprocess was clearly shown during period VII, when a higher syngas amount (26.2 mmol/L.day) was fedto both the reactors, while the organic loading was doubled in comparison to period VI. The methaneproduction in the RMBR was higher than the production in the FCBR (176.5 and 36.2 mL/L.day,respectively). These results indicate that the RMBR had the ability to maintain a high methanationcapacity, by retaining the microbial cells under the high risk of cell-wash out condition. Consequently,the system was able to convert more substrates into methane than the one with the free cells. Thisresult stresses the importance of cell-retention inside anaerobic digestion as shown in other studiesin which biofilm was used for high cell-density. However, with the method of cell-enclosure in themembranes, there is no danger of cell-detachment in contrast with the use of biofilm [26].

In the last period, the higher co-substrate concentration was fed to the systems in order toinvestigate if the product profiles changed. The results showed that the methane production in boththe reactors was maintained at a low level, observed as a lower utilization of H2 and CO (Figure 3a,b)and higher concentration of VFAs (Table 2). The co-substrate was not completely converted by themicrobial cells in this period. This result indicates that the maximum conversion capacity of thesystems had been reached. The lower methane production was probably because of the inhibitoryeffect of a high hydrogen partial pressure from the syngas, as well as the accumulation of the volatilefatty acids from the organic substances [2]. However, a higher amount of methane was still formed inthe RMBR compared to the FCBR. This is an effect of the higher amount of microbial cells retained inthe RMBR, still able to consume the co-substrates and convert them into methane.

Figure 3. Comparison of the performance of the reverse membrane bioreactor (RMBR) system andthe free cell bioreactor (FCBR) during the semi-continuous biomethanation of syngas and organicsubstances. The utilization of (a) Hydrogen; (b) Carbon monoxide; and (c) Carbon dioxide.

Page 57: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

Fermentation 2016, 2, 8 8 of 12

The above results indicate that during 154 days of syngas biomethanation, the RMBR systemwas an efficient and practical process in retaining the microbial cells, resulting in a high methaneproduction under a semi-continuous process condition.

3.2. Syngas Utilization Efficiency

Periods I, II, and III were designed to allow the microbial cells to adapt to the environmentalconditions in the beginning of the experiment. The artificial syngas fed was gradually increased, whilethe organic loading rate was controlled at a very low level (0.02 gCOD/L.day). The medium usedonly contained nutrients and vitamins during these periods. The flow rate of the gas recirculationwas 100 mL/min in the first experiment, and thereafter increased to 200 mL/min during periods IIand III in order to test the performances of the systems with shortened gas retention time. Figure 3and Table 3 present the percentage of gas utilization during the semi-continuous syngas fermentationprocess of 154 days.

Table 3. Investigated parameters during the semi-continuous biomethanation process ofthe co-substrates.

PeriodsSyngas Component Utilization (%) RMBR * Syngas Component Utilization (%) FCBR **

H2 CO CO2 H2 CO CO2

I 88.9 ˘ 0.0 80.8 ˘ 0.4 ´0.0 ˘ 0.5 96.5 ˘ 0.0 87.2 ˘ 0.2 ´1.7 ˘ 0.0II 100.0 ˘ 0.0 98.7 ˘ 0.0 ´16.6 ˘ 0.1 100.0 ˘ 0.0 99.5 ˘ 0.0 ´21.7 ˘ 0.0III 100.0 ˘ 0.0 100.0 ˘ 0.0 ´11.4 ˘ 0.2 100.0 ˘ 0.0 100.0 ˘ 0.0 ´12.2 ˘ 0.1IV 100.0 ˘ 0.0 98.2 ˘ 0.1 ´39.3 ˘ 0.1 100.0 ˘ 0.0 99.8 ˘ 0.0 ´36.3 ˘ 0.0V 99.7 ˘ 0.0 100.0 ˘ 0.0 ´17.6 ˘ 0.2 99.5 ˘ 0.0 100.0 ˘ 0.0 ´23.8 ˘ 0.2VI 94.7 ˘ 0.2 88.4 ˘ 0.7 ´40.3 ˘ 0.1 59.3 ˘ 0.6 42.3 ˘ 3.3 ´22.3 ˘ 0.1VII 56.6 ˘ 0.2 46.8 ˘ 4.9 25.6 ˘ 0.7 16.4 ˘ 0.3 38.9 ˘ 5.7 0.8 ˘ 0.7

* The reverse membrane bioreactor. ** The free cells bioreactor.

The amount of H2 in both the reactors remained at a very low level even when the H2 (syngas)fed was increased from 1.1 to 2.8 mmol/L.day, and the gas circulation rate was increased from 100 to200 mL/min. The percentage of H2 consumed in the RMBR and the FCBR was close to 100 percent.This shows that the microbial cells in both the RMBR and the FCBR could convert the H2 easily within1 day. Similar to the conversion of H2 in both the reactors during these periods, the microorganismsalso converted all the CO, without any observed negative effects. The CO in the gas fed was increasedfrom 3.1 to 7.7 mmol/L.day (Table 1), and it was utilized by the microbial cells in both the systemsat almost the same rate (about 100 percent). The CO2 amounts showed a different trend from theother gases. It increased during the first 3 days, and then stayed stable throughout periods I–III.It was observed that the CO2 amounts in both the reactors increased rather than decreased duringthese periods (a CO2 amount of 1.6–1.7 mmol/L.day was present in the syngas fed). This trendindicated that the microbial cells in both the reactors had performed as expected, with the CO2 beingboth produced and consumed during the fermentations. Thus, even though CO2 was supplementedtogether with the other compounds in the syngas, CO2 was also produced, leading to a stable amountof CO2 (Figure 3c). The most interesting conclusion that can be drawn from the results of these periodsis that the gases transfer through the PVDF membrane used, as the cell encasing material was not alimiting factor during the semi-continuous fermentation process, since the amount of the differentcomponents of the syngas were at similar levels as in the free cell system. The encased cells in theRMBR could assimilate the syngas continuously for product formation.

During periods IV and V, the gas circulation rate was increased to 300 mL/min, reducing the gasretention time in the systems. The syngas amount was decreased slightly while higher organic loadingrates with a medium containing acetic, propionic, and butyric acid was introduced. The results showedthat the syngas conversion performance of both the reactors was maintained during the simultaneous

Fermentation 2016, 2, 8 9 of 12

co-substrate fermentation. The microorganisms in both of the systems utilized almost 100 percent ofthe H2 and CO. However, CO2 started accumulating during these periods, indicating that the microbialcells decomposed the organic acids.

During the last two periods, VI and VII, a higher concentration of co-substrates was providedto the reactors. The results showed that the syngas utilization showed a decreasing trend in both thereactors, despite a higher concentration being added to the reactors during period VII (Table 1). Thehigher gas amounts, especially of H2, in the systems probably increased the partial pressure, therebyleading to a higher pH of the liquid solution (Figure 4b). Furthermore, additional organic substances,which resulted in the accumulated VFA, were present during these periods (Figure 4a). This resultedin a decreased performance of the microorganisms in the reactors. However, the RMBR still showed abetter ability to utilize the high concentration of co-substrates than the FCBR during these periods,shown by a higher percentage of H2 and CO utilization (Table 2). The percentage of H2 utilizationin the RMBR and the FCBR from period VI was 94.7 and 59.3, respectively. The CO utilization inthe RMBR and the FCBR was 88.4 and 42.3 percent, respectively. CO2 was accumulated, resultingin negative utilization values of ´40.3 and ´22.3 percent, respectively. These results revealed thatthe RMBR had a better capacity than the conventional system of the FCBR to convert higher syngasand co-substrate concentrations in a semi-continuous fermentation process. However, a positive CO2

utilization in both the reactors was observed in period VII. This means that CO2 was consumed ata higher rate than it was produced. In general, CO2 is produced during the anaerobic digestion oforganic compounds. In period VII, the VFAs were accumulated, meaning that the organic substanceswere not much used. This leads to less CO2 production in both the RMBR and FCBR. Although ahigher amount of CO2 was fed to the reactors (Table 1), CO2 was likely utilized by the microbial cells,thus resulting in a positive CO2 utilization. It can be concluded that retaining the microorganismsin the RMBR was a successful technique for improving the syngas biomethanation in the long-termfermentation process. Encased microbial cells in the RMBR showed a similar performance to assimilatethe syngas in a semi-continuous biomethanation process as the system with the free cells at a loworganic loading rate. Moreover, when higher syngas and organic loading rates were applied, a higherprocess efficiency of the RMBR, compared to the FCBR, was observed.

Figure 4. Accumulation of organic acids and pH during the semi-continuous biomethanation ofco-substrates. (a) Volatile fatty acids; (b) pH.

Page 58: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

Fermentation 2016, 2, 8 10 of 12

3.3. Volatile Fatty Acids and pH Profile

In this work, the performance of the bioreactors with cells degrading a mixture of two substratesin a semi-continuous process was investigated. Organic acids, including acetic, propionic, and butyricacid, were used as co-substrates to the syngas in the fermentations. The organic loading rate (OLR),in the form of the organic acids fed, was gradually increased in both the systems. The data oftotal volatile fatty acids (VFAs) in the effluents represented the organic acids utilization during thesimultaneous fermentation process. The VFA value is also an indicator for monitoring the stability ofthe anaerobic digestion process [4]. In a previous work, repeated batch syngas biomethanation wasexamined [27]. All of the medium was removed, and the encased cells were reused in a new batchexperiment. In that setup, the inhibiting substances probably disappeared, which is different from thecontinuous process of the present work, where most of the substrates remained in the reactors. Thus,the trend of the VFA content and the pH values were investigated in this experiment.

Figure 4 represents the total VFA concentration (a) and the pH values (b) in the effluents duringthe semi-continuous fermentation processes in the two investigated cell systems. The VFA amountsin both the reactors were close to zero during the periods I, II, III, and IV, even when the organicloading was increased from 0.02 to 0.43 g COD/L.day. This indicates that the microorganisms in boththe reactors were able to convert the organic substances simultaneously with syngas, without anyobserved negative effect. However, the pH value decreased slightly when the organic acids wereintroduced to the system during these periods. The pH was in the range of 7.0–7.8, and the changeswere not observed to affect the stability of the processes.

A low amount of VFAs in the reactors shows that the microorganisms are coping wellwith the process conditions. In the anaerobic digestion process, VFAs can be consumed by themethane-producing microorganisms. This reduction of VFAs increases the alkalinity of the surroundingmedium. The pH in the anaerobic digester is significantly affected by the production of VFAs and theCO2 content of the biogas.

In period V (days 88–126), there were substantially lower VFA concentrations measured in theeffluent from the RMBR (around 0.2 g/L) compared to the free cell system FCBR (around 0.6 g/L),as the OLR was increased to 0.86 gCOD/L.day. However, the VFAs concentration still remained at alow level during this period. The VFAs concentration in the RMBR was totally consumed. The reasonfor this fact was probably the high active microbial cell density retained in the RMBR, while the cellwash out was visually occurring from the FCBR. During this experiment, the pH value of both thereactors started to ramp up slightly. This probably happened because of the high conversion rate ofthe VFAs to biogas (methane and CO2), as well as the additional CO2 fed contained in the syngas(Figure 4b). In particular, the value from the RMBR had lower residual VFAs content than the FCBR.The pH increased to 8.7 and 8.1 in the RMBR and the FCBR, respectively.

When the concentration of the organic acids and syngas were increased in the last period, theaccumulation of the VFAs was even higher, while the pH dropped slightly. The VFA accumulation inthe FCBR increased, reaching a maximum amount of 9.4 g/L, while the total VFAs in the RMBR weredetected at 7.8 g/L, indicating that the cell encasement in the RMBR helped to maintain a higher abilityto degrade the VFAs compared to the FCBR in a semi-continuous fermentation process. However, therecommended VFA concentrations during the anaerobic digestion, without negatively affecting themethane-producing microorganisms in the process, are under 4 g/L [1,2]. At the observed levels in thereactors, it was thus clear that the maximum process capacity had been reached in this condition.

These results indicated that the RMBR system performed efficiently during the semi-continuoussyngas biomethanation with the addition of organic substances as a co-substrate. Organic acids weredegraded simultaneously with the reduction of the syngas until too high amounts of substrates werefed to the systems.

Fermentation 2016, 2, 8 11 of 12

4. Conclusions

The syngas biomethanation under simultaneous semi-continuous fermentation could be enhancedby retaining a high microbial cell density in a reverse membrane bioreactor (RMBR). No cell washoutwas visually observed from the RMBR during the process, running for 154 days. The performanceof the RMBR for the simultaneous fermentation of syngas and the organic compounds was betterthan the conventional system of free cells (FCBR). The microorganisms in the RMBR were able toconvert the syngas simultaneously with organic acids to methane at higher substrate concentrationsand productivity than the FCBR system. The RMBR reached maximum capacities of H2, CO,and CO2 conversion of 7.0, 15.2, and 4.0 mmol/L.day, respectively, at the organic loading rate of3.40 gCOD/L.day. The highest methane production rate from the RMBR was 186.0 mL/Lreactor.day onthe 147th day, compared to the highest rate in the FCBR, 106.3 mL/Lreactor.day, on the 58th day.

Acknowledgments: This work was financially supported by the Swedish Research Council. The authors wouldlike to thank Johan Westman for the useful discussions.

Author Contributions: All authors contributed to the conception and design of experiments; Supansa Y. Westmanand Konstantinos Chandolias performed the experiments and analyzed the data; all authors contributed to writingthe paper.

Conflicts of Interest: The authors declare no conflict of interest.

References

1. Deublein, D.; Steinhauser, A. Biogas from Waste and Renewable Resources; Wiley-VCH Verlag GmbH & Co.:Weinheim, Germany, 2008; pp. 49–50.

2. Gerardi, M.H. The Microbiology of Anaerobic Digesters; John Wiley & Sons: Hoboken, NJ, USA, 2003.3. Nizami, A.-S.; Korres, N.E.; Murphy, J.D. Review of the integrated process for the production of grass

biomethane. Environ. Sci. Technol. 2009, 43, 8496–8508. [CrossRef] [PubMed]4. Santosh, Y.; Sreekrishnan, T.R.; Kohli, S.; Rana, V. Enhancement of biogas production from solid substrates

using different techniques—A review. Bioresour. Technol. 2004, 95, 1–10.5. Daniell, J.; Köpke, M.; Simpson, S. Commercial biomass syngas fermentation. Energies 2012, 5, 5372–5417.

[CrossRef]6. Mohammadi, M.; Najafpour, G.D.; Younesi, H.; Lahijani, P.; Uzir, M.H.; Mohamed, A.R. Bioconversion

of synthesis gas to second generation biofuels: A review. Renew. Sust. Energy Rev. 2011, 15, 4255–4273.[CrossRef]

7. Munasinghe, P.C.; Khanal, S.K. Biomass-derived syngas fermentation into biofuels: Opportunities andchallenges. Bioresour. Technol. 2010, 101, 5013–5022. [CrossRef] [PubMed]

8. Kimmel, D.E.; Klasson, K.T.; Clausen, E.C.; Gaddy, J.L. Performance of trickle-bed bioreactors for convertingsynthesis gas to methane. Appl. Biochem. Micro. 1991, 28–29, 457–469. [CrossRef]

9. Henstra, A.M.; Sipma, J.; Rinzema, A.; Stams, A.J.M. Microbiology of synthesis gas fermentation for biofuelproduction. Curr. Opin. Biotech. 2007, 18, 200–206. [CrossRef] [PubMed]

10. Klasson, K.T.; Ackerson, M.D.; Clausen, E.C.; Gaddy, J.L. Bioconversion of synthesis gas into liquid orgaseous fuels. Enzym. Microb. Technol. 1992, 14, 602–608. [CrossRef]

11. Klasson, K.T.; Elmore, B.B.; Vega, J.L.; Ackerson, M.D.; Clausen, E.C.; Gaddy, J.L. Biological production ofliquid and gaseous fuels from synthesis gas. Appl. Biochem. Biotechnol. 1990, 24–25, 857–873. [CrossRef]

12. Lee, P.-H.; Ni, S.-Q.; Chang, S.-Y.; Sung, S.; Kim, S.-H. Enhancement of carbon monoxide mass transferusing an innovative external hollow fiber membrane (HFM) diffuser for syngas fermentation: Experimentalstudies and model development. Chem. Eng. J. 2012, 184, 268–277. [CrossRef]

13. Madigan, M.T.; Martinko, J.M.; Parker, J. Brock biology of Microorganisms; Prentice-Hall: Upper Saddle River,NJ, USA, 1997.

14. Orgill, J.J.; Atiyeh, H.K.; Devarapalli, M.; Phillips, J.R.; Lewis, R.S.; Huhnke, R.L. A comparison of masstransfer coefficients between trickle-bed, hollow fiber membrane and stirred tank reactors. Bioresour. Technol.2013, 133, 340–346. [CrossRef] [PubMed]

Page 59: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

Fermentation 2016, 2, 8 12 of 12

15. Younesi, H.; Najafpour, G.; Ku Ismail, K.S.; Mohamed, A.R.; Kamaruddin, A.H. Biohydrogen production in acontinuous stirred tank bioreactor from synthesis gas by anaerobic photosynthetic bacterium: Rhodopirillumrubrum. Bioresour. Technol. 2008, 99, 2612–2619. [CrossRef] [PubMed]

16. Shen, Y.; Brown, R.; Wen, Z. Enhancing mass transfer and ethanol production in syngas fermentationof clostridium carboxidivorans p7 through a monolithic biofilm reactor. Appl. Energy 2014, 136, 68–76.[CrossRef]

17. Bredwell, M.D.; Worden, R.M. Mass-transfer properties of microbubbles. 1. Experimental studies.Biotechnol. Prog. 1998, 14, 31–38. [CrossRef] [PubMed]

18. Shen, Y.; Brown, R.; Wen, Z. Syngas fermentation of clostridium carboxidivoran p7 in a hollow fibermembrane biofilm reactor: Evaluating the mass transfer coefficient and ethanol production performance.Biochem. Eng. J. 2014, 85, 21–29. [CrossRef]

19. Liu, K.; Atiyeh, H.K.; Stevenson, B.S.; Tanner, R.S.; Wilkins, M.R.; Huhnke, R.L. Continuous syngasfermentation for the production of ethanol, n-propanol and n-butanol. Bioresour. Technol. 2014, 151, 69–77.[CrossRef] [PubMed]

20. Wikandari, R.; Youngsukkasem, S.; Millati, R.; Taherzadeh, M.J. Performance of semi-continuous membranebioreactor in biogas production from toxic feedstock containing d-limonene. Bioresour. Technol. 2014, 170,350–355. [CrossRef] [PubMed]

21. Youngsukkasem, S.; Akinbomi, J.; Rakshit, S.K.; Taherzadeh, M.J. Biogas production by encased bacteria insynthetic membranes: Protective effects in toxic media and high loading rates. Environ. Technol. 2013, 34,2077–2084. [CrossRef] [PubMed]

22. Youngsukkasem, S.; Barghi, H.; Rakshit, S.; Taherzadeh, M. Rapid biogas production by compact multi-layermembrane bioreactor: Efficiency of synthetic polymeric membranes. Energies 2013, 6, 6211–6224. [CrossRef]

23. Youngsukkasem, S.; Rakshit, K.S.; Taherzadeh, M.J. Biogas production by encapsulated methane producingbacteria. Bioresources 2012, 7, 56–65.

24. Osuna, M.B.; Zandvoort, M.H.; Iza, J.M.; Lettinga, G.; Lens, P.N. Effects of trace element addition on volatilefatty acid conversions in anaerobic granular sludge reactors. Environ. Technol. 2003, 24, 573–587. [CrossRef][PubMed]

25. Isci, A.; Demirer, G.N. Biogas production potential from cotton wastes. Renew. Energy 2007, 32, 750–757.[CrossRef]

26. Langer, S.; Schropp, D.; Bengelsdorf, F.R.; Othman, M.; Kazda, M. Dynamics of biofilm formation duringanaerobic digestion of organic waste. Anaerobe 2014, 29, 44–51. [CrossRef] [PubMed]

27. Youngsukkasem, S.; Chandolias, K.; Taherzadeh, M.J. Rapid bio-methanation of syngas in a reversemembrane bioreactor: Membrane encased microorganisms. Bioresour. Technol. 2015, 178, 334–340. [CrossRef][PubMed]

© 2016 by the authors; licensee MDPI, Basel, Switzerland. This article is an open accessarticle distributed under the terms and conditions of the Creative Commons by Attribution(CC-BY) license (http://creativecommons.org/licenses/by/4.0/).

76

76

Paper III

Effects of heavy metals and pH on the conversion of biomass to hydrogen via syngas

fermentation

Page 60: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

PEER-REVIEWED ARTICLE bioresources.com

Chandolias et al. (2018). “Biomass to hydrogen,” BioResources 13(2), 4455-4469. 4455

Effects of Heavy Metals and pH on the Conversion of Biomass to Hydrogen via Syngas Fermentation Konstantinos Chandolias,a,* Steven Wainaina,a Claes Niklasson,b and Mohammad J. Taherzadeh a

The effects of three heavy metals on hydrogen production via syngas fermentation were investigated within a metal concentration range of 0 to 1.5 mg Cu/L, 0 to 9 mg Zn/L, 0 to 42 mg Mn/L, in media with initial pH of 5, 6, and 7, at 55 °C. The results showed that at lower metal concentration, pH 6 was optimum while at higher metal concentrations, pH 5 stimulated the process. More specifically, the highest hydrogen production activity recorded was 155% ± 12% at a metal concentration of 0.04 mg Cu/L, 0.25 mg Zn/L, and 1.06 mg Mn/L and an initial medium pH of 6. At higher metal concentration (0.625 mg Cu/L, 3.75 mg Zn/L, and 17.5 mg Mn/L), only pH 5 was stimulating for the cells. The results showed that the addition of heavy metals, contained in gasification-derived ash, can improve the production rate and yield of fermentative hydrogen. This could lead to lower costs in gasification process and fermentative hydrogen production and less demand for syngas cleaning before syngas fermentation.

Keywords: Gasification; Syngas; Fermentative hydrogen; Heavy metals; pH Contact information: a: Swedish Centre for Resource Recovery, University of Borås, S-50190 Borås, Sweden; b: Department of Chemistry and Chemical Engineering, Chalmers University of Technology, S-412 96 Gothenburg, Sweden; *Corresponding author: [email protected] INTRODUCTION

To face the increasing global waste generation and energy demand, sustainable strategies for the conversion of waste into green energy are required. Anaerobic fermentation is a relatively simple and cost-effective method that converts organic wastes, such as food residuals, sewage waste, and manure, into biofuels. However, there is a big fraction of solid waste, such as forest residues and mixed landfill wastes, which are difficult or impossible to degrade biologically because of their complex structure. An effective way to treat this recalcitrant biomass is with the combination of gasification and anaerobic fermentation processes. During this two-stage process, the feedstock is converted into syngas, a gas consisting mainly of hydrogen, carbon monoxide, and carbon dioxide. Thereafter, anaerobic bacteria digest the syngas and create value-added chemicals, such as volatile fatty acids (VFA), and biofuels such as hydrogen.

Hydrogen is a valuable gas with a high-energy content of 120 MJ/kg to 142 MJ/kg, which is higher than that of hydrocarbon fuels by a factor of approximately 2.75 and releases only water during combustion (Henstra et al. 2007; Tufa et al. 2016). This gas has numerous applications in chemical processes, lamps, balloons, vehicle fuels, laboratories, as a reductive agent (redox reactions), etc. Hydrogen is also a product of several processes. The largest fraction of the global hydrogen production (80% to 85%) is obtained by the steam methane reforming (SMR) process of natural gas (Fan et al. 2016), which uses fossil

Page 61: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

PEER-REVIEWED ARTICLE bioresources.com

Chandolias et al. (2018). “Biomass to hydrogen,” BioResources 13(2), 4455-4469. 4456

fuels as the main substrate. In addition, electricity can be stored in hydrogen form through the power-to-gas process. Another process for hydrogen generation, with less environmental drawbacks but higher costs, is the water electrolysis method, although it currently can only be used in small-scale applications (Tufa et al. 2016). Biocatalysts, such as microalgae and phototrophic bacteria, have been employed in special-designed photobioreactors for hydrogen production (Skjånes et al. 2016). Another biological process is the dark fermentation of biomass such as food residues, manure, and straw. The main advantage of the fermentation process for hydrogen production is that it is less expensive, less energy-demanding, and more eco-friendly than the other conventional processes (Levin et al. 2004; Lin and Shei 2008). All the above processes can lead to a substantial production of hydrogen to be used for energy and as material resources.

The anaerobic syngas conversion into hydrogen can occur via a biological water gas shift process. In this process, carbon monoxide acts as an electron donor and the CO-dehydrogenase (CODH) enzyme provides electrons and protons for the carbon monoxide. Thereafter, the electrons released by the oxidation reaction are transferred to another enzyme called energy-converting hydrogenase (ECH) (Henstra et al. 2007), which catalyzes the reduction of electrons for the production of molecular hydrogen. Fatty acids are by-products that can be produced through the H2/CO2 or CO consumption pathway (Grimalt-Alemany et al. 2017), shown in Eq. 1,

CO + H2O → CO2 + H2, ΔG0 = -20 KJ/mol (1)

At the end of the gasification process, inorganic components, such as heavy metals, still remain in the residual ashes (Dong et al. 2015). These metals can be present in different forms such as metal oxides or metal chlorides (Liao et al. 2007; Tafur-Marinos et al. 2014), and their composition depends on the feedstock and gasification conditions (Dong et al. 2015). Despite the fact that the gasification ashes have been used as building materials for roads and other constructions, they can leach and pose a threat to the environment and to the public’s health (James et al. 2012). The main threat derives from the heavy metals, which are toxic and classified as known or possible carcinogens for humans, according to the U.S. Environmental Protection Agency and the International Agency for Research on Cancer (Tchounwou et al. 2012). In contrast, heavy metals have been shown to be beneficial for the growth of fermentative microbes (Osuna et al. 2003). For instance, a study on fermentative hydrogen production reported that the addition of Cu and Cr ions improved the hydrogen yield (Lin and Shei 2008). Therefore, the use of heavy metals during syngas fermentation could reduce the environmental footprint of gasification while enhancing the fermentation process.

Although several studies have been conducted on the effects of heavy metals on anaerobic methane production (Fang and Chan 1997; Lin and Chen 1999; Abdel-Shafy and Mansour 2014), there are only a handful of studies on the effects of heavy metals on the fermentative hydrogen production (Li and Fang 2007; Lin and Shei 2008). According to the authors’ best knowledge, this is the first study that investigates the impact of heavy metals in syngas fermentation. The aim of this work is to provide valuable data on the beneficial and inhibitory effects of heavy metals and pH during syngas fermentation and suggest strategies for improving the efficacy of this process. This knowledge can be used in order to reduce the environmental impact of gasification and reduce the costs of syngas cleaning and fermentative hydrogen production.

PEER-REVIEWED ARTICLE bioresources.com

Chandolias et al. (2018). “Biomass to hydrogen,” BioResources 13(2), 4455-4469. 4457

EXPERIMENTAL Materials Liquid medium

The liquid medium inside the reactors consisted of micronutrients in trace amounts, macronutrients, and the three heavy metals that were under investigation (except the control reactors).

The concentration of the inorganic macronutrients in the used liquid medium was 280 mg NH4Cl/L, 330 mg K2HPO4•3H2O/L, 100 mg MgSO4•7H2O/L, and 10 mg CaCl2•2H2O/L (Osuna et al. 2003). Moreover, 1 mL of the trace element stock solution was contained in 1 L of the liquid medium. The concentration of the trace element stock solution included 2000 mg FeCl2•4H2O/L, 50 mg H3BO3/L, 50 mg ZnCl2/L, 500 mg MnCl2•4H2O/L, 38 mg CuCl2•2H2O/L, 50 mg (NH4)6Mo7O24•4H2O/L, 2000 mg CoCl2•6H2O/L, 142 mg NiCl2•6H2O/L, and 164 mg Na2SeO3•5H2O/L (Osuna et al. 2003).

The metal ions used were in the form of metal oxides (ZnO, CuO) and salt (MnCl2•4H2O) powder. Initially, 0.25 g of each metal powder was dissolved in 20 mL of 2% nitric acid (HNO3), and deionized water was added to achieve a total volume of 500 mL. The metal stock solutions were dissolved to achieve a metal ratio (Cu:Zn:Mn) of approximately 1:6:28. This metal ratio is similar to the heavy metal composition that was found in the fly and bottom ashes of a wood pellet pyro-gasification facility (Tafur-Marinos et al. 2014).

Syngas composition

A commercial syngas mixture containing carbon monoxide (55 mol%), hydrogen (20 mol%), and carbon dioxide (10 mol%) (Klasson et al. 1990) was provided by AGA Gas AB (Gothenburg, Sweden). The overpressure in the gas cylinder containing the syngas was built by nitrogen gas. Inoculum

The mixed anaerobic consortium was obtained from a local thermophilic 3000 m3 anaerobic digester that typically digested the organic fraction of municipal solid waste (Borås Energy and Environment, Borås, Sweden). A typical pH value in the digester was 8 to 8.2 and the main fraction of the substrate was food waste. The total solids (% TS) and volatile solids (% VS) of the inoculum were 14.23% ± 0.27% and 14.15% ± 0.27%, respectively. Methods Reactor characteristics, inoculation, and start-up

The reactors were serum glass bottles with plastic caps, rubber sealing, and a total volume of 118 mL (Bioprocess Control AB, Lund, Sweden). The anaerobic culture was incubated at 55 °C for 5 days to consume all the nutrients prior to the experiment. After incubation, the excess water from the inoculum was removed by centrifugation at 4300 ×g for 10 min. Then, each reactor was loaded with 3 g of inoculum and filled with 40 mL of liquid medium. The headspace of each reactor was purged with syngas for 3 min with a flow rate of 50 mL/min in order to remove air and feed the reactor with the gaseous substrate. After the syngas-purging, the pressure inside the headspace was adjusted at 1

Page 62: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

PEER-REVIEWED ARTICLE bioresources.com

Chandolias et al. (2018). “Biomass to hydrogen,” BioResources 13(2), 4455-4469. 4458

atm with a needle that removed the excess gas. The reactors were continuously shaken (100 rpm) in a warm water bath (55 °C ± 1 °C) and placed at a 45° angle. The gaseous samples were collected from the top of the reactors through the rubber sealing regularly, and the liquid samples were collected at the end of the experiment from the liquid medium.

The inoculum was acclimated to syngas as the sole carbon and energy source for 15 days before the experimental results were obtained. The experiment was conducted in batch mode and in triplicates, and the results are presented as the mean values ± standard deviation. The duration of the experiment was 96 h. Gas analysis

Gas samples were collected from the headspace of the bioreactors every 12 h with a 0.25-mL gas-tight syringe (VICI, Precision Sampling Inc., Baton Rouge, LA, USA). The gas components were analyzed by using gas chromatography (GC). For the analysis of hydrogen and carbon dioxide, a Perkin-Elmer gas chromatograph (Clarus 500; Norwalk, CT, USA) was used, equipped with a packed column (CarboxenTM 1000, 6’ × 1.8” OD, 60/80 Mesh, Supelco, Shelton, CT, USA) using a thermal conductivity detector (Perkin-Elmer, Norwalk, CT, USA) with an injection temperature of 200 °C. The levels of the carbon monoxide were also analyzed by another gas chromatograph (Clarus 400; Perkin-Elmer, Norwalk, CT, USA), equipped with a packed column (CarboxenTM 1000, 6’ × 1.8” OD, 60/80 Mesh, Supelco, Shelton, CT, USA) and a thermal conductivity detector (Perkin-Elmer, Norwalk, CT, USA) with an injection temperature of 200 °C. Both of the gas chromatographs used nitrogen as a carrier gas with a flow rate of 30 mL/min at 75 °C.

Liquid analysis

The VFA content in the liquid samples was analyzed using a Waters 2695 high-performance liquid chromatograph (HPLC; Waters Corporation, Milford, CT, USA) with a hydrogen-based column (Aminex HPX87-H; BioRad Laboratories, München, Germany) at 60 °C and 0.6 mL/min (5 mM H2SO4 eluent) equipped with a refractive index (RI) detector (Waters 2410, Waters Corporation, Milford, CT, USA). The metal concentrations of the liquid samples were measured with a microwave plasma-atomic emission spectrometer (MP-AES 4200; Agilent Technologies, Santa Clara, CA, USA) equipped with an inlet air filter kit - 4107 nitrogen generator (Agilent Technologies, Santa Clara, CA, USA) and an SPS 3 Autosampler (Agilent Technologies, Santa Clara, CA, USA). RESULTS AND DISCUSSION

The coupling of thermochemical and biochemical processes, such as gasification and dark fermentation, can convert recalcitrant biomass into hydrogen. The use of heavy metals derived from the ashes of thermochemical processes can reduce the environmental footprint of these processes and boost the efficacy of biochemical processes.

This work focused on the biochemical conversion of syngas into hydrogen through dark fermentation. The goal was to investigate the effect of heavy metals and pH and to report the optimum, inhibiting, and toxic metal concentrations and pH values. Hydrogen production activity (Eq. 2), hydrogen yield (Eq. 3), and total VFA yield (Eq. 4) were employed to monitor the extent of heavy metal and pH effects on the syngas fermentation

PEER-REVIEWED ARTICLE bioresources.com

Chandolias et al. (2018). “Biomass to hydrogen,” BioResources 13(2), 4455-4469. 4459

process. The heavy metal uptake (%), was also obtained. Equation 2 shows the hydrogen production activity (%),

AH (%) = (Ai / Ac) × 100 (2)

where Ac is the hydrogen amount (mmol) produced by the control reactors, and Ai is the hydrogen amount (mmol) produced by the metal-dosed reactors after 96 h of anaerobic fermentation. The yield values were calculated as follows,

YH = H2 produced / CO fed (3) YVFA = VFA concentration / CO fed (4)

where YH is the hydrogen yield (mmol H2/mmol CO fed) and YVFA is the VFA yield (g VFA/L/mmol CO fed) obtained after 96 h of anaerobic fermentation. The Effect of HNO3

The addition of HNO3 in the reactors had two main purposes. First, the acid dissolved the heavy metal compounds inside the liquid medium and, second, the addition of HNO3 shifted the biological activity from methane to hydrogen production. Thus, no inoculum pretreatment that favored hydrogen production and the inhibition of methanogens was necessary. Figure 1 (panels a through d) shows how the consumption/production of gas components was affected by the addition of HNO3. The HNO3 concentration that inhibited methane production was 0.08%. This nitrate concentration was relatively low in comparison to the concentration used in other studies. For example, in a work that investigated the effect of ammonia and nitrate on biogas production from food waste, no inhibitory effect on methane production was reported at total ammonia nitrate addition levels below 1.1 g/L (Sheng et al. 2013). However, these inhibitory levels depend also on the anaerobic species of the inoculum.

Several mechanisms have been proposed to explain the inhibition of methanogens by HNO3. One possible mechanism is that nitrate raises the redox potential, Eh, of the medium (Jones 1972). Another reason may be the toxic effects of nitrogen intermediates such as nitrite and nitrous oxide (Iwamoto et al. 1999; Ungerfeld 2015). Recent studies suggest that nitrate can be used as an alternative hydrogen sink so that it can compete with available hydrogen, which is one of the main methane substrates together with carbon dioxide (Yang et al. 2016). However, in this work, hydrogen production was favored, and therefore it can be assumed that the inhibiting mechanism was based on the toxicity of nitrate on methanogens.

The Effect of Heavy Metals and pH

Syngas fermentation for hydrogen production took place in bioreactors containing media with heavy metal concentrations between 0 mg Cu/L, 0 mg Zn/L, 0 mg Mn/L, and 1.5 mg Cu/L, 9 mg Zn/L, 42 mg Mn/L and the different initial medium pHs that were investigated were 7, 6, and 5. The metals were added in a specific ratio in the medium to mimic the real metal composition in the ash of a wood pellet gasification plant as described previously.

Page 63: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

PEER-REVIEWED ARTICLE bioresources.com

Chandolias et al. (2018). “Biomass to hydrogen,” BioResources 13(2), 4455-4469. 4460

Fig. 1. Effect of the addition of HNO3 during syngas fermentation

PEER-REVIEWED ARTICLE bioresources.com

Chandolias et al. (2018). “Biomass to hydrogen,” BioResources 13(2), 4455-4469. 4461

Figure 2 shows the relationships between the accumulated hydrogen production, heavy metal concentrations, and initial pH of the medium.

Fig. 2. Relationships between accumulative hydrogen production, initial medium pH of 7 (a), 6 (b), and 7 (c), and heavy metal concentrations (mg/L)

Page 64: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

PEER-REVIEWED ARTICLE bioresources.com

Chandolias et al. (2018). “Biomass to hydrogen,” BioResources 13(2), 4455-4469. 4462

The accumulative hydrogen amount was calculated by the total biogas production and hydrogen gas content. Figure 2 also shows the time (h) that was required for each bioreactor to reach its maximum hydrogen production in 96 h of fermentation. The results showed that the different heavy metal loadings and the initial pHs of the medium influenced the hydrogen production.

The role of heavy metals in anaerobic fermentation is essential, as many of them are part of the enzymes as metallic co-factors. More specifically, Mn stabilizes the enzyme methyltransferase (MPB) and affects redox reactions (Fisher et al. 1973; Perry and Silver 1982), Cu affects the superoxide dismutase (SODM) and hydrogenase in MPB in Clostridia and facultative anaerobes (Jones et al. 1987; Kirby et al. 1981), and Zn affects the activity of hydrogenase in MPB, the sulfate-reducing bacteria (SRB), and formate dehydrogenase (FDH) (Adams et al. 1986; Kirby et al. 1981). A total lack of or overdose of heavy metals may cause the inhibition or loss of enzyme functions. Usually, at high concentrations, metals act as nonspecific, reversible inhibitors and do not compete with the substrate. In this type of inhibition, the metals bind to enzymes and form the enzyme-inhibitor (EI) or enzyme-substrate-inhibitor (ESI) complexes of EI or ESI type (Wood and Wang 1983). However, some heavy metals, such as Zn and Cd, may cause competitive inhibition, meaning that they may compete with the substrate. Possible microbial inhibition mechanisms include substitution of the metallic enzyme co-factors, combining with the sulfhydryl group (-SH), and tight binding to acid groups in the polypeptide chains (Wood and Wang 1983).

The hydrogen production activity (AH) and hydrogen yield (YH) are shown in Fig. 3. The reactors dosed with low heavy metal concentrations showed a higher hydrogen production in comparison to the control reactors. In particular, the metal concentrations of 0.04 mg Cu/L, 0.25 mg Zn/L, and 1.06 mg Mn/L and 0.1 mg Cu/L, 0.67 mg Zn/L, and 2.8 mg Mn/L stimulated the hydrogen production activity (Ah > 100%). The highest hydrogen production activity recorded was 155% ± 12% at a metal concentration of 0.04 mg Cu/L, 0.25 mg Zn/L, and 1.06 mg Mn/L and initial medium pH 6. Another study reported that a metal concentration of 3 mg Cu/L resulted in a 10% to 20% increase in hydrogen production (Lin and Shei 2008). Other studies on fermentative hydrogen production also have shown that the highest hydrogen production was achieved in mediums with a pH of 6 and that lowering the pH inhibited hydrogen formation (Dareioti et al. 2014). This phenomenon depends on the type of the dominant bacterial species inside the anaerobic microbial community of the inoculum. For example, Clostridia, which are usually dominant in this process, function within a pH range of 6.0 to 6.7. The pH is a very crucial parameter for the microbial growth of anaerobic communities and it directly affects the fermentative hydrogen production. The pH of the medium changes the intracellular pH and the electrochemical gradient across the microbial membrane and thus influences the regulation of metabolism and the bioenergetics of microorganisms.

Metal concentrations higher than 0.1 mg Cu/L, 0.67 mg Zn/L, and 2.8 mg Mn/L were demonstrated to be inhibitory for the cells in media with pH 6 and 7 (Fig. 3). More specifically, media containing metal mixtures of 1.25 mg Cu/L, 7.57 mg Zn/L, and 35 mg Mn/L and 1.5 mg Cu/L, 9 mg Zn/L, and 42 mg Mn/L were shown to be inhibitory and reduced the hydrogen production activity to values as low as 35%. Another study reported a 50% reduction in the hydrogen production activity for mixed inoculum that was dosed with media containing 4.5 mg Zn/L and 6.5 mg Cu/L (Lin and Shei 2008). These

PEER-REVIEWED ARTICLE bioresources.com

Chandolias et al. (2018). “Biomass to hydrogen,” BioResources 13(2), 4455-4469. 4463

concentrations are higher but comparable to the inhibiting metal concentrations found in this study. However, the metal concentrations of 0.625 mg Cu/L, 3.75 mg Zn/L, and 17.5 mg Mn/L stimulated the hydrogen production activity in mediums with a pH of 5. Therefore, for substrates with higher metal concentrations, pH 5 is recommended. This shift in the preference of pH at higher metal concentrations is possibly connected to the types of different strains of the microbial community and their optimum conditions of growth.

Fig. 3. Relationships between hydrogen production activity, Ah (%) = (mol H2/mol H2 control)*100 (a), hydrogen yield, YH = mol H2 prod./mol CO fed (b), initial medium pH and heavy metal concentrations (mg/L)

Page 65: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

PEER-REVIEWED ARTICLE bioresources.com

Chandolias et al. (2018). “Biomass to hydrogen,” BioResources 13(2), 4455-4469. 4464

The hydrogen yield showed a similar result to the hydrogen production activity (Fig. 3). The hydrogen yields were stimulated at lower metal concentrations and inhibited at higher metal concentrations. The highest hydrogen yield achieved was 1.32 mol ± 0.02 mol H2 prod./mol CO fed in bioreactors containing 0.1 mg Cu/L, 0.67 mg Zn/L, and 2.8 mg Mn/L and medium with initial pH 7.

Volatile fatty acid production was detected only in the control reactors (Fig. 4). This meant that the presence of heavy metals inhibited complete acidogenesis even though hydrogen formation was favored at low metal concentrations (Fig. 3). The VFA formation increased as pH decreased and the highest VFA yield reported was 1.086 g VFA/L per mmol of carbon monoxide fed, at a pH of 5.

Fig. 4. Relationships between VFA yield, YVFA = gVFA/L/mmol CO fed, initial medium pH, and heavy metal concentrations (mg/L) Heavy Metal Uptake (%)

During anaerobic fermentation, the heavy metals can end up inside the cells or they can be absorbed or precipitated with other inorganics (Hayes and Theis 1978; Howgrave-Graham and Wallis 1991). The availability of heavy metals as nutrients or as inhibitors is affected by the concentration of the metals as well as other parameters such as the pH of the liquid medium (Oleszkiewicz and Sharma 1990). However, even if there is a full availability of the metals as nutrients, this does not mean that the microorganisms utilize the metals. The metal uptake may be hindered by the presence of other metals, the lack of carrier molecules inside the cells, high excretion of metal ions by the cells, and the failure of the energy driven system (Wood and Wang 1983). In the case of the opposite conditions, excessive metal uptake may take place. In general, the uptake of heavy metals and all other nutrients indicate the growth rate of the cells (Oleszkiewicz and Sharma 1990).

PEER-REVIEWED ARTICLE bioresources.com

Chandolias et al. (2018). “Biomass to hydrogen,” BioResources 13(2), 4455-4469. 4465

Fig. 5. Relationships between metal uptake (%) of Zn (a) and Mn (b), initial medium pH, and initial heavy metal concentrations (mg/L)

Page 66: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

PEER-REVIEWED ARTICLE bioresources.com

Chandolias et al. (2018). “Biomass to hydrogen,” BioResources 13(2), 4455-4469. 4466

In this work, the metal uptake was calculated from the initial and final metal concentrations inside the liquid medium of the reactors (Fig. 5). The results showed that the uptake of Zn and Mn was approximately 100% up to a metal concentration of 0.1 mg Cu/L, 0.67 mg Zn/L, and 2.8 mg Mn/L. At a metal concentration of 0.625 mg Cu/L, 3.75 mg Zn/L, and 17.5 mg Mn/L the metal uptake decreased to 90 80%, while at a higher metal concentration of 1.25 mg Cu/L, 7.57 mg Zn/L, and 35 mg Mn/L the Zn and Mn uptake dropped especially in media with lower pHs. This result was similar to the findings of another study that investigated the uptake of several heavy metals, at different pHs, during anaerobic digestion (Wang et al. 2007). That study also concluded that heavy metal uptake increased at higher pH values. In addition, several studies have shown that low pHs may hinder the heavy metal uptake (Cheng et al. 1975; Wang et al. 1999) because of the pH value’s influence on the binding of metals on the microbial cell wall (Wang et al. 2007). However, the pH inhibition level also depended on the type of each metal. In studies performed with activated sludge, it was reported that the main binding sites of the metals on the sludge sites were amino acids on the cell wall. Some metals, such as Cu and Pb, exhibit a strong affinity for anaerobic sludge sites (Artola et al. 1997), while other metals, such as Zn and Ni, show a relatively low affinity (Leighton and Forster 1997; Artola et al. 2000).

The above results showed that the addition of heavy metals during syngas fermentation improved hydrogen production. In addition, the appropriate combination of metal concentrations and initial pH of the liquid medium increased the efficacy of the process. CONCLUSIONS 1. The addition of Cu, Zn, and Mn improved hydrogen production during syngas

fermentation.

2. The highest hydrogen production activity achieved was 155% ± 12% at a metal concentration of 0.04 mg Cu/L, 0.25 mg Zn/L, and 1.06 mg Mn/L and a pH of 6.

3. At a higher metal concentration of 0.625 mg Cu/L, 3.75 mg Zn/L, and 17.5 mg Mn/L, pH 5 was optimum.

4. The uptake of Zn and Mn decreased at higher metal doses and lower pHs. ACKNOWLEDGMENTS

The authors gratefully acknowledge the financial support of the Swedish Research Council.

PEER-REVIEWED ARTICLE bioresources.com

Chandolias et al. (2018). “Biomass to hydrogen,” BioResources 13(2), 4455-4469. 4467

REFERENCES CITED Abdel-Shafy, H. I., and Mansour, M. S. M. (2014). "Biogas production as affected by

heavy metals in the anaerobic digestion of sludge," Egypt. J. Petrol. 23(4), 409-417. DOI: 10.1016/j.ejpe.2014.09.009

Adams, M. W. W., Jin, S.-L. C., Chen, J.-S., and Mortenson, L. E. (1986). "The redox properties and activation of the F420)-non-reactive hydrogenase of Methanobacterium formicicum,” BBA- Protein Struct. M. 869(1), 37-47. DOI: 10.1016/0167-4838(86)90307-9

Artola, A., Balaguer, M. D., and Rigola, M. (1997). "Heavy metal binding to anaerobic sludge,” Water Res. 31(5), 997-1004. DOI: 10.1016/S0043-1354(96)00345-4

Artola, A., Martin, M., Balaguer, M., and Rigola, M. (2000). "Isotherm model analysis for the adsorption of Cd (II), Cu (II), Ni (II), and Zn (II) on anaerobically digested sludge,” J. Colloid Interf. Sci. 232(1), 64-70. DOI: 10.1006/jcis.2000.7186

Cheng, M. H., Patterson, J. W., and Minear, R. A. (1975). "Heavy metals uptake by activated sludge,” Water Pollut. Control Fed. 47(2), 362-376.

Dareioti, M. A., Vavouraki, A. I., and Kornaros, M. (2014). "Effect of pH on the anaerobic acidogenesis of agroindustrial wastewaters for maximization of bio-hydrogen production: A lab-scale evaluation using batch tests,” Bioresource Technol. 162, 218-227. DOI: 10.1016/j.biortech.2014.03.149

Dong, J., Chi, Y., Tang, Y. J., Ni, M. J., Nzihou, A., Weiss-Hortala, E., and Huang, Q. X. (2015). "Partitioning of heavy metals in municipal solid waste pyrolysis, gasification, and incineration,” Energ. Fuel. 29(11), 7516-7525. DOI: 10.1021/acs.energyfuels.5b01918

Fan, J. M., Zhu, L., Jiang, P., Li, L. L., and Liu, H. M. (2016). "Comparative exergy analysis of chemical looping combustion thermally coupled and conventional steam methane reforming for hydrogen production,” J. Clean. Prod. 131, 247-258. DOI: 10.1016/j.jclepro.2016.05.040

Fang, H. H. P., and Chan, O. C. (1997). "Toxicity of electroplating metals on benzoate-degrading granules,” Environ. Technol. 18(1), 93-99. DOI: 10.1080/09593331808616516

Fisher, S., Buxbaum, L., Toth, K., Eisenstadt, E., and Silver, S. (1973). "Regulation of manganese accumulation and exchange in Bacillus subtilis W23,” J. Bacteriol. 113(3), 1373-1380.

Grimalt-Alemany, A., Skiadas, I. V., and Gavala, H. N. (2017). "Syngas biomethanation: State-of-the-art review and perspectives,” Biofuel. Bioprod. Bior. 12(1), 139-158. DOI: 10.1002/bbb.1826

Hayes, T. D., and Theis, T. L. (1978). “The distribution of heavy metals in anaerobic digestion,” Water Pollut. Control Fed. 50(1), 61-72.

Henstra, A. M., Sipma, J., Rinzema, A., and Stams, A. J. M. (2007). "Microbiology of synthesis gas fermentation for biofuel production,” Curr. Opin. Biotech. 18(3), 200-206. DOI: 10.1016/j.copbio.2007.03.008

Howgrave-Graham, A. R., and Wallis, F. M. (1991). "Preparation techniques for the electron microscopy of granular sludge from an anaerobic digester,” Lett. Appl. Microbiol. 13(2), 87-89. DOI: 10.1111/j.1472-765X.1991

Page 67: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

PEER-REVIEWED ARTICLE bioresources.com

Chandolias et al. (2018). “Biomass to hydrogen,” BioResources 13(2), 4455-4469. 4468

Iwamoto, M., Asanuma, N., and Hino, T. (1999). "Effects of nitrate combined with fumarate on methanogenesis, fermentation, and cellulose digestion by mixed ruminal microbes in vitro,” Nihon Chikusan Gakk. 70(6), 471-478. DOI: 10.2508/chikusan.70.471

James, A. K., Thring, R. W., Helle, S., and Ghuman, H. S. (2012). "Ash management review-applications of biomass bottom ash,” Energies 5(10), 3856-3873. DOI: 10.3390/en5103856

Jones, G. A. (1972). "Dissimilatory metabolism of nitrate by the rumen microbiota,” Can. J. Microbiol. 18(12), 1783-1787. DOI: 10.1139/m72-279

Jones, W. J., Nagle, D. P., and Whitman, W. B. (1987). "Methanogens and the diversity of archaebacteria,” Microbiol. Rev. 51(1), 135-177.

Kirby, T. W., Lancaster, J. R., and Fridovich, I. (1981). "Isolation and characterization of the iron-containing superoxide dismutase of Methanobacterium bryantii,” Arch. Biochem. Biophys. 210(1), 140-148. DOI: 10.1016/0003-9861(81)90174-0

Klasson, K. T., Elmore, B. B., Vega, J. L., Ackerson, M. D., Clausen, E. C., and Gaddy, J. L. (1990). "Biological production of liquid and gaseous fuels from synthesis gas,” Appl. Biochem. Biotech. 24(1), 857-873. DOI: 10.1007/BF02920300

Leighton, I. R., and Forster, C. F. (1997). "The adsorption of heavy metals in an acidogenic thermophilic anaerobic reactor,” Water Res. 31(12), 2969-2972. DOI: 10.1016/S0043-1354(97)00159-0

Levin, D. B., Pitt, L., and Love, M. (2004). "Biohydrogen production: Prospects and limitations to practical application,” Int. J. Hydrogen Energ. 29(2), 173-185. DOI: 10.1016/S0360-3199(03)00094-6

Li, C., and Fang, H. H. P. (2007). "Inhibition of heavy metals on fermentative hydrogen production by granular sludge,” Chemosphere 67(4), 668-673. DOI: 10.1016/j.chemosphere.2006.11.005

Liao, C., Wu, C., and Yan, Y. (2007). "The characteristics of inorganic elements in ashes from a 1 MW CFB biomass gasification power generation plant,” Fuel Process. Technol. 88(2), 149-156. DOI: 10.1016/j.fuproc.2005.06.008

Lin, C.-Y., and Chen, C.-C. (1999). "Effect of heavy metals on the methanogenic UASB granule,” Water Res. 33(2), 409-416. DOI: 10.1016/S0043-1354(98)00211-5

Lin, C.-Y., and Shei, S.-H. (2008). "Heavy metal effects on fermentative hydrogen production using natural mixed microflora,” Int. J. Hydrogen Energ. 33(2), 587-593. DOI: 10.1016/j.ijhydene.2007.09.030

Oleszkiewicz, J. A., and Sharma, V. K. (1990). "Stimulation and inhibition of anaerobic processes by heavy metals—A review,” Biol. Waste. 31(1), 45-67. DOI: 10.1016/0269-7483(90)90043-R

Osuna, M. B., Zandvoort, M. H., Iza, J. M., Lettinga, G., and Lens, P. N. (2003). "Effects of trace element addition on volatile fatty acid conversions in anaerobic granular sludge reactors,” Environ. Technol. 24(5), 573-587. DOI: 10.1080/09593330309385592

Perry, R. D., and Silver, S. (1982). "Cadmium and manganese transport in Staphylococcus aureus membrane vesicles,” J. Bacteriol. 150(2), 973-976.

Sheng, K., Chen, X., Pan, J., Kloss, R., Wei, Y., and Ying, Y. (2013). "Effect of ammonia and nitrate on biogas production from food waste via anaerobic digestion," Biosyst. Eng. 116(2), 205-212. DOI: 10.1016/j.biosystemseng.2013.08.005

PEER-REVIEWED ARTICLE bioresources.com

Chandolias et al. (2018). “Biomass to hydrogen,” BioResources 13(2), 4455-4469. 4469

Skjånes, K., Andersen, U., Heidorn, T., and Borgvang, S. A. (2016). "Design and construction of a photobioreactor for hydrogen production, including status in the field,” J. Appl. Phycol. 28(4), 2205-2223. DOI: 10.1007/s10811-016-0789-4

Tafur-Marinos, J. A., Ginepro, M., Pastero, L., Torazzo, A., Paschetta, E., Fabbri, D., and Zelano, V. (2014). "Comparison of inorganic constituents in bottom and fly residues from pelletised wood pyro-gasification,” Fuel 119, 157-162. DOI: 10.1016/j.fuel.2013.11.042

Tchounwou, P. B., Yedjou, C. G., Patlolla, A. K., and Sutton, D. J. (2012). "Heavy metals toxicity and the environment,” Molecular, Clinical and Environmental Toxicology 101, 133-164. DOI: 10.1007/978-3-7643-8340-4_6

Tufa, R. A., Rugiero, E., Chanda, D., Hnàt, J., Van Baak, W., Veerman, J., Fontananova, E., di Profio, G., Drioli, E., Bouzek, K., et al. (2016). "Salinity gradient power-reverse electrodialysis and alkaline polymer electrolyte water electrolysis for hydrogen production,” J. Membrane Sci. 514, 155-164. DOI: 10.1016/j.memsci.2016.04.067

Ungerfeld, E. M. (2015). "Shifts in metabolic hydrogen sinks in the methanogenesis-inhibited ruminal fermentation: A meta-analysis,” Front. Microbiol. 6(37). DOI: 10.3389/fmicb.2015.00037

Wang, J. M., Huang, C. P., and Allen, H. E. (2007). "Effect of pH on metal uptake by anaerobic sludge,” Environ. Eng. Sci. 24(8), 1095-1104. DOI: 10.1089/ees.2006.0267

Wang, J. M., Huang, C. P., Allen, H. E., Poesponegoro, I., Poesponegoro, H., and Takiyama, L. R. (1999). "Effects of dissolved organic matter and pH on heavy metal uptake by sludge particulates exemplified by copper(II) and nickel(II): Three-variable model,” Water Environ. Res. 71(2), 139-147. DOI: 10.2175/106143099X121517

Wood, J. M., and Wang, H. K. (1983). "Microbial resistance to heavy metals,” Environ. Sci. Technol. 17(12), 582A-590A. DOI: 10.1021/es00118a002

Yang, C., Rooke, J. A., Cabeza, I., and Wallace, R. J. (2016). "Nitrate and inhibition of ruminal methanogenesis: Microbial ecology, obstacles, and opportunities for lowering methane emissions from ruminant livestock,” Front. Microbiol. 7, 132. DOI: 10.3389/fmicb.2016.00132

Article submitted: January 9, 2018; Peer review completed: April 9, 2018; Revised version received: April 19, 2018; Accepted: April 25, 2018; Published: April 30, 2018. DOI: 10.15376/biores.13.2.4455-4469

Page 68: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

77

Paper IV

Protective effect of RMBR against syngas impurities

Page 69: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

1

Protective Effect of a Reverse Membrane Bioreactor against

Toluene and Naphthalene in Anaerobic Digestion

Konstantinos Chandoliasa, Laurenz Alan Ricardo Sugiantob, Nurina Izazib, Ria Millatib*,

Rachma Wikandarib, Päivi Ylitervoa, Claes Niklassonc, Mohammad Taherzadeha

a Swedish Center for Resource Recovery, University of Borås, 50190 Borås, Sweden,

bDepartment of Food and Agricultural Product Technology, Universitas Gadjah Mada,

Yogyakarta 55281, Indonesia,

cDepartment of Chemistry and Chemical Engineering, Chalmers University of Technology,

41296 Gothenburg, Sweden

*corresponding author: [email protected]

Page 70: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

2

Abstract

A combined route for biogas production is thermochemical conversion of wastes to syngas

(H2, CO and CO2) followed by their biological conversion into CH4 during anaerobic

digestion. However, raw syngas contains tar contaminants that can be toxic for the cells.

This work investigated the effect of two common tar contaminants, toluene and

naphthalene, at a concentration range of 0 6.44 g/L and 0 1.28 g/L, respectively, on

methane production, with a mixed culture. Free cell bioreactors and reverse membrane

bioreactors were operated in parallel, during batch and continuous thermophilic anaerobic

digestion. The results from the continuous digestion showed inhibition at a toluene and

naphthalene concentration of 3.14 g/L and 0.63 g/L, respectively in the free cell

bioreactors. At a toluene and naphthalene concentration of 4.80 g/L and 0.96 g/L, the CH4

production rate in membrane bioreactors was approximately 47% and 160% higher than in

the free cell bioreactors.

Keywords: Anaerobic Digestion; Syngas Contaminants; Reverse Membrane Bioreactor;

Naphthalene; Toluene; Protective Effect.

3

1. Introduction

Biofuel production is one of the main strategies of international policies, such as

Agenda 21 and Kyoto Protocol, in order to meet the increasing energy demand, the waste

generation and the environmental pollution caused by the use of fossil fuels (Weiland,

2010). Biogas, containing mainly CH4 and CO2, is produced during anaerobic digestion of

organic material and it is considered as a sustainable alternative to fossil fuels. The use of

upgraded biogas (>90% methane), in transportation sector can cause a significant

reduction in greenhouse gas emissions, in comparison to gasoline (Sahota et al., 2018).

Biogas production creates also decentralized energy production sites. This valuable gas

product is used for heat and electricity production, as vehicle fuel, cooking or it can be

injected to the natural gas grid (Pöschl et al., 2010). However, there are limitations to the

type of substrate that the anaerobic microorganisms can degrade.

Poorly degradable and non-degradable substrates are recalcitrant and require a

pretreatment prior to their anaerobic digestion. One such pretreatment is gasification, a

thermochemical process in which the feedstock is thermally degraded at high temperature

and at the presence of a controlled amount of an oxidizing agent, such as O2, air or steam.

In comparison to other pretreatment processes, such as enzymatic and acid hydrolysis,

gasification gathers advantages, such as feedstock flexibility, high conversion rates and

rapid mass and volume waste reduction (Consonni et al., 2005; Gershman, 2013). The

product of gasification of carbonaceous feedstock is syngas, a mixture of mainly CO, H2

and CO2. This versatile gas has numerous applications, such as electricity generation and

production of methanol, ammonia, mixed alcohols, acids, waxes, olefins, and biofuels

(Spath & Dayton, 2003).

Page 71: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

4

The Fischer-Tropsch process is widely used in industrial application for syngas

conversion into biofuels, via the use of metal catalysts. Prior to this process, syngas has to

be upgraded, by removing contaminants. Typical syngas contaminants are particulate

matters, tars, NOx, S, N2, and alkali (Woolcock & Brown, 2013). Tars are condensable

organic compounds, which are mainly formed due to the operating parameters of

gasification. The type of tar contaminants varies from primary to heavier oxygenated

hydrocarbons and polycyclic aromatic hydrocarbons (Stevens, 2001). Several syngas-

cleaning methods are available, such as hot gas clean-up (HGC), inertial separation

(cyclons) and barrier filtration (Woolcock & Brown, 2013), however, these methods have

a relatively high cost (Patinvoh et al., 2017).

During the past years, the biological conversion of syngas into methane and value-

added chemicals via anaerobic fermentation has gained increasing interest in scientific,

social, and industrial fields (Bengelsdorf et al., 2018). In comparison to catalytic syngas

conversion, such as Fischer-Tropsch process, syngas fermentation is more product

specific, less sensitive to CO/H2/CO ratio, generates no hazardous components and does

not require high temperature and pressure (Drzyzga et al., 2015). Furthermore, syngas can

be co-fermented with other carbonaceous substrates, such as food waste, in order to

increase the fermentation efficiency. In a study of syngas fermentation, the addition of

synthetic organic waste containing acetic, butyric and propionic acid, caused an increase

in CH4 production (Youngsukkasem et al., 2015).

One of the main challenges in syngas biomethanation is the low gas-to-liquid mass

transfer, which is greatly affected by the low gas hold up in the bioreactors. A recent study

5

reported higher syngas conversion rates in a reverse membrane bioreactor (RMBR) due to

a higher gas hold up (Chandolias et al., 2019). This type of membrane bioreactor contains

anaerobic cells enclosed inside polyvinylidene difluoride (PVDF) membrane sachets that

float inside the medium of the bioreactor. The same type of bioreactor was used in order to

protect the anaerobic cells against fruit flavours, such as D-limonene in another study

(Wikandari et al., 2014). The hydrophilic surface of the membranes blocked successfully

the hydrophobic inhibitor, while volatile fatty acids were able to completely permeate the

membranes after 30 min. The RMBR has also been used in order to avoid the cell washout

and achieve high cell-density during continuous syngas fermentation (Westman

Youngsukkasem et al., 2016).

The RMBR could also be used in order to protect the anaerobic cells from the

contaminants present in the raw syngas. This would eliminate the costly requirement for

syngas cleaning prior to syngas fermentation. Syngas contaminants can inhibit the

microbial performance or change other parameters such as pH and redox potential

(Grimalt-Alemany et al., 2017). Toluene and naphthalene are two common chemical

compounds present in tars that can be found in syngas (Lopes et al., 2018). These

compounds can be used as a microbial substrate at low amounts (Maillacheruvu & Pathan,

2009), however, high concentrations could be toxic for the cells (Edwards & Grbić-Galić,

1994; Foght, 2008; Marozava et al., 2018). Although studies have shown that low

amounts of syngas contaminants do not affect the anaerobic syngas conversion, there is no

sufficient knowledge on the minimum gas-cleaning requirements before biomethanation.

Moreover, different anaerobic microorganisms have different tolerance towards syngas

contaminants. Therefore, further investigation is required in order to study the inhibitory

Page 72: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

6

effect of syngas contaminants in the conversion of syngas into biogas (Grimalt-Alemany

et al., 2017). The most sensitive anaerobic microbes are the methanogens and therefore a

strong and robust methanogen community is required. This can be achieved by feeding the

cells with adequate carbon and energy source, such as a synthetic organic medium

(Youngsukkasem et al., 2015).

In this work, the effect of toluene and naphthalene was investigated during anaerobic

fermentation of synthetic organic medium by a mixed culture, at thermophilic conditions.

In addition, a RMBR was employed in order to study a possible protective effect of the

hydrophilic membranes against high concentrations of the hydrophobic toluene and

naphthalene. The experiments were operated in both batch and continuous conditions and

the RMBRs were operated in parallel with free cell bioreactors (FCBR).

2. Materials and Methods

2.1 Inoculum and membrane-encasement

The inoculum used was a mixed culture collected from a recirculating stream of a local

3000-m3 anaerobic digester (Borås Energy and Environment AB, Borås, Sweden), which

operated at thermophilic conditions and was fed with food residues. Prior to the

experiment, the culture was incubated for 3 days at 55 °C so that all nutrients would be

consumed. After incubation, large particles and other debris were trapped by a 1-mm sieve

and the excess water was removed by centrifugation at 9000×g for 3 min (Hereaus

Megafuge 8, Thermo Scientific, Germany). The inoculum used had a total solid (TS) and

volatile solid (VS) content of 71.16% and 18.63%, respectively. The inoculum was placed

inside the reactors either as free (suspended) cells or as membrane-encased cells.

7

The membrane used was a hydrophilic polyvinylidene fluoride flat sheet membrane

with a pore size of 0.1μm (Durapore, Thermo Fisher Scientific Inc., Sweden). The cell-

encasement process has been described in detail in a previous study (Youngsukkasem et

al., 2012).

2.2 Liquid medium

The liquid medium used in this work was a mixture stock solution. The stock solution

contained basal medium (macronutrients and micronutrients) and organic acids (Osuna et

al., 2003). Toluene and naphthalene dissolved in methanol were mixed with the stock

solution before the feeding of the bioreactors. Methanol was selected as a solvent because

at low concentrations, it has no inhibitory effect for the cells and it is a good solvent of

toluene and naphthalene, which are not easily dissolved in water solutions (Dickhut et al.,

1989; Du et al., 2016). In the stock solution, the concentration of the macronutrients was

28 mg/L NH4Cl, 33 mg/L K2HPO43H2O, 10 mg/L MgSO47H2O, and 1 mg/L

CaCl22H2O. The concentration of the micronutrients was 2 mg/L FeCl24H2O, 0.5 mg/L

H3BO3, 0.5 mg/L ZnCl2, 5 mg/L MnCl24H2O, 0.038 mg/L CuCl22H2O, 0.05 mg/L

(NH4)6Mo7O244H2O, 2 mg/L CoCl26H2O, 0.142 mg/L NiCl26H2O, and 0.164 mg/L

Na2SeO35H2O. The concentration of the acids was 10 g/L acetic acid, 3.33 g/L propionic

acid and 3.33 g/L butyric acid. The toluene and naphthalene solutions had a concentration

of 234.4 g/L and 46.88 g/L, respectively. Different volumes of these solutions were added

into the bioreactors in order to achieve toluene and naphthalene concentrations of 0 6.44

g/L and 0 2.46 g/L, respectively. The methanol concentration in the feed was ≤ 0.034%.

The pH of the medium fed to the bioreactors was adjusted to 7.0 ± 0.2 with 5M NaOH.

Page 73: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

8

2.3 Reactor characteristics, seeding and experimental setup

The batch bioreactors were serum glass bottles (Bioprocess control AB, Lund, Sweden)

with a total volume of 118 mL. Each bioreactor was loaded with 23.34 mL liquid medium

and seeded with 3 g of suspended cells, in FCBRs or a membrane sachet containing 3 g or

encased cells, in RMBRs. The batch digestion was run in duplicates and the batch

bioreactors were continuously shaken in a water batch (55 °C), at an inclination of

approximately 45 ° and a shaking frequency of 100 rpm.

The continuous bioreactors were glass bottles (Bioprocess control AB, Lund, Sweden)

with a total volume of 600 mL. Each bioreactor was loaded with 350 mL liquid medium

and seeded with 45 g of suspended cells, in FCBRs, or equal amount of encased cells in

RMBRs (3 g/membrane). The continuous bioreactors were placed in a shaking water bath,

at 55 °C and 45 rpm, without inclination.

After seeding, the bioreactors were sealed with butyl caps and then purged with N2, in

order to remove oxygen from the bioreactor´s headspace. The ratio of mass of inoculum to

volume of liquid substrate was approximately 0.13 g/mL in both batch and continuous

bioreactors. This ratio has an direct effect on the product yields as it has been reported in a

previous study (Chandolias et al., 2019).

2.4 Analytical methods

During the experiment, the composition of the gas headspace of the bioreactors and the

liquid medium were regularly analysed, in order to ensure the good operation of the

bioreactors. In the continuous experiments, the outlet of the bioreactors was connected to

an Automatic Methane Potential Test System (AMPTS II, Bioprocess control AB, Lund,

Sweden), which measured the volume of the daily gas production. Gas samples were

9

collected from the headspace of the bioreactors and analysed with gas chromatography

(GC). The gas samples (100 μL) were collected with a 0.25-mL gas-tight syringe (VICI,

Precision Sampling Inc., Baton Rouge, LA, USA) and injected into a Varian 450 Gas

Chromatograph (Palo Alto, CA, USA). The GC was equipped with a wall coated open

tubular (WCOT) capillary column (WCOT, J&W Scientific GC-Gas Pro, bonded silica

based 30 m × 0.32 mm, Agilent Technologies, Santa Clara, CA, USA) and a thermal

conductivity detector (TCD, Varian, Palo Alto, CA, USA) and Galaxie Chromatography

Data System Single Instrument (v.1.9, Varian, Walnut Creek, CA, USA) as the analysis

software. The GC was connected to a computer using Galaxie Chromatography Data

System v.1.9 Single Instrument as the data recording software. N2 was used as carrier gas

with a flow rate of 2.0 mL/min, through the column and 30 mL/min, through the detector.

The temperatures of injector, oven, detector, heater, and detector filament were set to 75

°C, 60 °C, 120 °C, and 200 °C, respectively, and the injection split ratio was set to 5.

The amount of toluene and naphthalene in the bioreactors was measured with a gas

chromatograph with mass spectroscopy detector, GC-MS (GC-MS Trace GC Ultra,

Themoscientific, Waltham, MA, USA) equipped with a capillary column (Zebron ZB-5ms

fused silica, 30 m×0.25 mm×0.25 μm, Torrance, CA, USA). The injection had a split of

50:1, temperature of 230 °C and 1 μL of sample. Helium was used as carrier gas at a

constant flow of 0.7 mL/min and the oven was set to 40 °C for 2 min and 210 °C for 3 min

with a ramp of 17 °C/min, while the MS detector was set to 250 °C. The preparation of the

liquid samples was done with and without solvent extraction. For the solvent extraction, 1

mL of sample was mixed with 1 mL of diethyl ether and then shaken overnight, at room

temperature. Sonication of 5 min was used in order to disrupt the cells.

Page 74: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

10

3. Results and Discussion

Syngas components are good substrate for CH4 production during anaerobic

fermentation. However, syngas contaminants may have an inhibitory effect on anaerobic

microorganisms. Thus, cleaning of raw syngas prior to syngas fermentation is required.

However, the cost of this gas-cleaning is significant in comparison to the overall operating

cost of the process. In addition, few information is available on the inhibitory effect of

various syngas contaminants.

This work investigated the effect of two common tar contaminants in raw syngas,

toluene and naphthalene, during the anaerobic digestion of synthetic organic substrate and

the cell-encasement into membrane sachets inside a RMBR as a protective mean. Each

RMBR was operated in parallel with a FCBR, containing free/suspended cells. The

experiments took place in both batch and continuous conditions while the accumulative

CH4 production, CH4 production rate, and CH4 production activity were used as indicators

of the bioreactors` efficacy. The results showed good tolerance of the cells against high

concentrations of toluene and naphthalene compared to previous studies and improvement

of the CH4 production rate in RMBRs, in which toluene and naphthalene were added.

3.1 Investigation of toluene and naphthalene effect in batch anaerobic digestion

A batch anaerobic digestion took place for 34 days and duplicate bioreactors were

operated with a synthetic organic medium with a COD content of 42 g COD/L. Prior to

the experiment, the mixed culture was acclimatized in synthetic organic medium for 35

days. The range of toluene and naphthalene concentrations investigated was 0 1.0 g/L and

0 0.2 g/L, respectively. These concentrations were higher than other inhibitory

concentrations found in the literature (Akyol et al., 2015; Dou et al., 2009). Toluene and

11

naphthalene are aromatic hydrocarbons, labelled as highly toxic to living organisms.

Toluene consists of one benzene ring and one methyl group, while naphthalene consists of

two benzene rings. The presence of toluene and naphthalene in anaerobic environments

can inhibit the cellular metabolism. Naphthalene can disrupt the membrane of anaerobic

cells and produce toxic metabolites. More specifically, the lipophilic nature of naphthalene

can alter the fluidity of the cell membrane, cause swelling of the lipid bilayer and affect

the intracellular energy transduction. The produced toxic metabolites can be more harmful

to the cells because of the disrupted cellular membrane, the production of reactive oxygen,

and the disturbance in DNA and protein formation (Pumphrey & Madsen, 2007). The

tolerance of anaerobic microorganisms to these compounds depends mainly on the type of

culture, compound concentrations, and type of substrate (Akyol et al., 2015; Dou et al.,

2009; Govind et al., 1991; Shimp & Pfaender, 1985). Studies in anaerobic digestion have

shown inhibitory effect of toluene and naphthalene at concentrations above 0.09 g/L and

0.01 g/L, respectively (Edwards & Grbić-Galić, 1994; Ince et al., 2009; Pumphrey &

Madsen, 2007).

Fig. 1 presents the total amount of CH4 obtained in bioreactors fed with 0.5 g/L and 1.0

g/L of toluene and 0.1 g/L and 0.2 g/L of naphthalene. The results showed no inhibitory

effect of toluene and naphthalene at the investigated concentrations in FCBRs. On the

contrary, CH4 production was increased at higher contaminant concentrations. More

specifically, the increase in CH4 production at a high toluene concentration of 1 g/L,

suggests that the cells were able to utilize this compound as a substrate (Fig 1A). Another

study investigated the effect of toluene on acetyl-CoA synthetase expression level of

acetoclastic methanogen, Methanosaeta concilii. The results showed complete inhibition

Page 75: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

12

after the 3rd exposure at toluene concentrations higher than 0.046 g/L, or 0.023 g toluene

per g VS inoculum (Akyol et al., 2015). In the current work, at toluene concentrations of

up to 1.0 g/L or at a ratio of 0.06 g toluene per g VS inoculum, no inhibition effect was

observed, however, the cells were not exposed to toluene multiple times.

Fig. 1B showed also a tendency for higher CH4 production at a high concentration of

0.2 g/L naphthalene or 0.012 g naphthalene per g VS inoculum. Another study on the

anaerobic degradation of naphthalene by mixed bacteria under nitrate reducing conditions,

showed also no inhibition and complete degradation of naphthalene within a period of 25

days incubation when the initial concentration was below 0.030 g/L (Dou et al., 2009).

Moreover, higher degradation rates of naphthalene were recorded in bioreactors with

higher initial concentrations of naphthalene (Dou et al., 2009). However, another work,

reported inhibition at lower naphthalene concentration. More specifically, a study on the

naphthalene metabolism and growth inhibition by naphthalene in Polaromonas

naphthalenivorans strain CJ2 showed inhibition for naphthalene concentrations above

0.01 g/L (Pumphrey & Madsen, 2007). In addition, naphthalene has been reportedly toxic

to the archetypal naphthalene degraders P. putida G7 and P. putida NCIB 9816-4 under

nitrogen or oxygen-limiting conditions (Ahn et al., 1998) or at high concentration of 2 g/L

of naphthalene crystals (Park et al., 2004).

The results shown in Fig. 1 did not indicate any inhibitory effect while CH4 production

was higher at higher contaminant concentration. The yields obtained were approximately

40 50% of the theoretical yield. The use of mixed consortia and the incubation with

organic acids prior to the experiment possibly made the mixed culture more tolerant to

toluene and naphthalene in comparison to other works. Previous studies have reported a

13

relation between the degradation of easily degradable substrates such as organic acids and

the effect of phenolic compounds during anaerobic digestion. For example, the cellular

adaptation to increasing concentrations of amino acids, carbohydrates, or fatty acids

improved the ability of aquatic bacteria to tolerate and degrade phenols (Shimp &

Pfaender, 1985). A mixed culture can be tolerant towards inhibitors due to the diversity of

microorganisms and the domination of the strongest cells according to the type of

substrate and growth conditions. Moreover, the tolerance at high concentrations of toluene

and naphthalene in this work could be a result of the relatively high amount of g VS

inoculum seeded in the bioreactors.

3.2 Effect of toluene and naphthalene in continuous anaerobic digestion

The results from batch experiment suggested that in order to observe inhibition in CH4

production, higher toluene and naphthalene concentrations should be investigated.

However, from the literature, different studies suggest different inhibiting concentrations

for these compounds depending on the type of culture, substrate, medium and other

parameters (Akyol et al., 2015; Edwards & Grbić-Galić, 1994; Ince et al., 2009).

Therefore, the contaminant concentrations were increased daily until inhibition on CH4

production was observed. A synthetic organic medium was used as a carbon and energy

source throughout the experiment; with an OLR of 1.24 g COD/Ld and a hydraulic

retention time of 17.5 days. During the feeding of FCBRs, cell washout was avoided by

centrifugation of the digestate and recycling of the pelleted cells into the bioreactors. Until

day-35, only synthetic organic medium was fed into the bioreactors in order to create a

strong and robust anaerobic culture. From day-35 until day-75, the concentrations of

Page 76: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

14

contaminants in the bioreactors were increased daily until inhibition was observed. From

day-75 until day-80, no contaminants were added into the bioreactors and from day-80

until day-120 the concentration of contaminants into the bioreactors was stable. More

specifically, from day-35 until day-55, 0.035 g of toluene and 0.007 g of naphthalene were

added daily into the bioreactors. From day-55 until day-75, 0.1755 g of toluene and 0.035

g of naphthalene were fed into the bioreactors daily. Between day-75 and day-80, no

contaminants were added into the bioreactors and from day-80 until day-120, 0.096 g of

toluene and 0.019 g of naphthalene were fed into the bioreactors daily. Fig. 2-4 present the

CH4 production rate, CH4 production activity, and the theoretical concentration of

contaminants inside the bioreactors, with the assumption that they were not consumed,

during the continuous experiment. The CH4 yield obtained in this study, in the bioreactor

with 0 g/L of contaminants was close to the theoretical yield (result not shown), which

indicated favorable conditions for CH4 production, during the anaerobic digestion. The

theoretical yield in the above calculations is based on the assumption that all organic

compounds were converted into CH4.

Fig. 2 shows the effect of different toluene concentrations on CH4 production rate and

production activity. The increase in CH4 production rate in FCBR, from day-35 until day-

55, shows that the anaerobic cells were able to grow and used toluene as a substrate.

Toluene is in general the most readily degraded compound of the monoaromatic BTEX

(benzene, toluene, ethylbenzene and xylene isomers) components. In comparison to

naphthalene, toluene can be degraded by a larger variety of anaerobic bacteria, which

belong mostly either to the Azoarcus or Thauera genus (Foght, 2008; Weelink et al.,

2010). Moreover, toluene degradation takes place in the presence of various terminal

15

electron acceptors such as nitrate, iron (III), sulfate and CO2 (Edwards & Grbić-Galić,

1994). In the presence of CO2, toluene can be directly converted into CH4 while

naphthalene can be converted into acetate and H2 (Christensen et al., 2004; Edwards &

Grbić-Galić, 1994). A study on anaerobic degradation by methanogenic consortium

showed that, after a long adaptation period of 100 days, the cultures were able to degrade

toluene but not naphthalene and that toluene degradation was inhibited by the presence of

acetate, propionate and methanol (Edwards & Grbić-Galić, 1994). However, in this

current work, the addition of toluene had a positive effect on the CH4 production rate.

Another study, on the acetyl-coA synthetase expression level of acetoclastic methanogen

Methanosaeta concilii showed positive effect during the co-digestion of toluene and

methanol. More specifically, the co-digestion of methanol and toluene did not cause

combined inhibition, however, singular toluene-exposed cells were inhibited (Akyol et al.,

2015). Therefore, the small amount of methanol used for dissolving naphthalene and

toluene in this work could even have had a positive effect on the degradation of these

compounds.

The CH4 production rate started to decrease on day-57 at a toluene concentration of

2.05 g/L (Fig. 2A) in the FCBR, while toluene inhibition was observed on day-60 at a

toluene concentration of 3.14 g/L when the CH4 production activity was lower than 100%

(Fig. 2B). The FCBR system showed signs of recovery during the no-feeding period and

during the last period with a stable toluene concentration of 4.8 g/L. The average CH4

production activity, in the last period, was approximately 100%. The inhibiting toluene

concentration obtained in the current work is quite high in comparison to the reports from

other similar studies. For example, during an anaerobic digestion with aquifer-derived

Page 77: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

16

mixed consortium, the observed inhibiting toluene concentration was higher than 0.092

g/L, after 100 120 days of adaptation (Edwards & Grbić-Galić, 1994). Another study with

anaerobic sludge in an up flow anaerobic sludge blanket bioreactor (UASB) reported a

62% inhibition on potential methane production (PMP) at a toluene concentration of 0.091

g/L (Ince et al., 2009). The system recovery phenomenon has been observed in other

studies with toluene and other aromatic hydrocarbons as substrates. Warthmann et al.

(2004) observed that xylene degradation could be inhibited by adding toluene to an

anaerobic sludge digester and it could be recovered again when toluene was not added.

Moreover, Datar et al. (2004) observed also that cells remained in the dormant state after

exposure to syngas contaminants and cells would start growing again after the

contaminant feeding was omitted.

The use of RMBRs could have a protective effect against syngas contaminants. RMBRs

have been used in previous studies in order to improve the gas hold up and consequently

the gas-to-liquid mass transfer in syngas fermentation (Chandolias et al., 2019) and as

protective barriers against fruit flavors during anaerobic digestion (Wikandari et al.,

2014). The membrane blocked the hydrophobic inhibitors because of its hydrophilic

surface while all substrate and products were free to pass through its pores. The RMBR

could also block the passage of hydrophobic inhibitors, such as tars, NOx, and particles,

which are present in the raw syngas after gasification.

In order to improve the cellular tolerance to substrates containing high concentrations of

toluene, a RMBR was operated in parallel with the FCBR. The increase in CH4 production

rate (RMBR) after toluene feeding on day-36 indicates that an amount of toluene could

pass through the membrane pores and that it was used as substrate by the cells. This result

17

agrees with the results from batch anaerobic digestion in the previous section. In contrast

to the FCBR, the CH4 production rate in RMBR increased from day-36 until day-67. This

proves that the membranes protected the cells from high toluene concentration while at the

same time; a fraction of toluene could enter the membranes and was converted to CH4. On

the other side, the free cells were inhibited due to direct exposure to the extreme toluene

concentrations. The stimulation of CH4 production in RMBR can be also verified by its

CH4 production activity, which was above 100%, during these days. After day-68, at a

toluene concentration of 5.25 g/L, the CH4 production rate decreased although the CH4

production activity remained above 100%. This indicates that the concentration of toluene

in the medium was too high even for the encased cells. The lowest CH4 production

activity observed in the RMBR was approximately 106%, at a toluene concentration of 5.4

g/L, on day-78, while the highest CH4 production activity was approximately 144% and it

was achieved at a stable toluene concentration of 4.8 g/L. This value was significantly

higher compared to the CH4 production activity in the FCBR of approximately 99% at the

same toluene concentration.

Naphthalene is considered as a more toxic component to the cells in comparison to

toluene, mainly due to the toxic metabolites that are generated during its degradation. In

comparison to toluene degrading microorganisms, fewer cells can degrade naphthalene. In

addition, the cells can use fewer electron acceptors in the presence of naphthalene

compared to toluene metabolism while both toluene and naphthalene degradation can

occur with nitrate as electron acceptor (Cunningham et al., 2001; Edwards & Grbic-Galic,

1992; Heider et al., 1998; Rockne & Strand, 2001). In a study, where an anaerobic culture

was enriched with 1-methylnaphthalene as a sole carbon source and Fe(OH)3 as electron

Page 78: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

18

acceptor, the gram-positive Thermoanaerobacteraceae could degrade polycyclic aromatic

hydrocarbons while the Desulfobacterales were mainly responsible for Fe(III) reduction

(Marozava et al., 2018). Because of the high toxicity of naphthalene, lower amount of this

inhibitor was added into the bioreactors in comparison to toluene.

Fig. 3 shows the effect of different naphthalene concentrations on the CH4 production

rate and production activity. The CH4 production rate in FCBR increased until day-57. On

day-60, inhibition was observed at a naphthalene concentration of 0.63 g/L. The methane

production increased when naphthalene feeding was paused between day-76 and day-80.

However, the CH4 production activity dropped from approximately 114% on day-80 to

41% on day-120, which indicates inhibition at a stable naphthalene concentration of 0.96

g/L. In another study on naphthalene inhibition during anaerobic digestion, the inhibitory

naphthalene concentration was approximately 0.01 g/L (Pumphrey & Madsen, 2007),

which is approximately 100-fold lower compared to the current results in this work. The

CH4 production in RMBR followed a similar trend with the FCBR until day-80. Between

day-80 until day-120, the highest CH4 production rate and CH4 production activity of

approximately 56.7 mmol/(Ld) and 175%, respectively, were achieved in the RMBR. In

addition, the CH4 production rate was rather stable during this period. This was probably a

result of cell adaptation at high naphthalene concentrations and proved the protective

effect of the cell encasement towards high naphthalene concentrations.

The effect of simultaneous toluene and naphthalene addition was examined and the

results are shown in Fig. 4. In the FCBR, inhibition occurred on day-60, at a toluene and

naphthalene concentration of 3.14 g/L and 0.63 g/L, respectively. After a short increase in

19

CH4 production rate between day-75 and day-80, the CH4 production activity dropped

from approximately 141% (day-80) to 28% (day-120), indicating inhibition. Interestingly

enough, no inhibition was observed in the RMBR at the highest toluene and naphthalene

concentrations of 6.44 g/L and 1.26 g/L, respectively. However, similar results were

obtained in both FCBR and RMBR between day-80 and day-120 due to severe cell

leaking from the membrane sachets in RMBR. From the comparison of Fig. 2 4, between

day-55 and day-75, it can be concluded that the presence of toluene had a positive effect

on anaerobic digestion with naphthalene also present in RMBRs. This was probably a

result of the greater diversity of microorganisms that thrived in the bioreactors containing

both toluene and naphthalene and created a more robust culture (Liang & Ramesh, 2012).

Some amount of the contaminants could have been accumulated inside the cells. This

explanation is verified by the result of liquid analysis that showed low amounts of

contaminants detected after cell disruption. The complete degradation of toluene and

naphthalene is considered not possible because of the high contaminant concentration.

Furthermore, the yields were low, approximately 40% of the theoretical yield in the

RMBR fed with naphthalene and less than 20% of the theoretical yield in all other

bioreactors fed with contaminants (after day-80). The inhibiting effect of the contaminants

was also depicted on the consumption of the organic medium since there was higher

accumulation of organic acids in the FCBRs especially from day-80 until day-120 in

comparison to the RMBRs (results not shown).

4. Conclusions

Batch and continuous anaerobic digestion took place in this study, for the investigation

of CH4 production in the presence of two syngas contaminants, toluene and naphthalene.

Page 79: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

20

The results showed no inhibition on CH4 production at low concentrations of toluene and

naphthalene. However, at higher concentrations of these contaminants, the CH4 production

was inhibited. In addition, significantly higher CH4 production rate and stimulation were

observed in RMBRs even at high concentrations of contaminants, in comparison to the

FCBRs.

Acknowledgements

The authors would like to express sincere gratitude to the Swedish Research Council and

Swedish Council for Higher Education for the financial support.

The authors declare no conflict of interest.

5. References

1. Ahn, I.S., Ghiorse, W.C., Lion, L.W., Shuler, M.L. 1998. Growth kinetics of Pseudomonas putida G7 on naphthalene and occurrence of naphthalene toxicity during nutrient deprivation. Biotechnol Bioeng, 59(5), 587-94, doi:10.1002/(SICI)1097-0290(19980905)59:53.0.CO;2-6.

2. Akyol, Ç., Ince, O., Coban, H., Koksel, G., Cetecioglu, Z., Ayman Oz, N., Ince, B. 2015. Individual and combined inhibitory effects of methanol and toluene on acetyl-CoA synthetase expression level of acetoclastic methanogen, Methanosaeta concilii. Int Biodeterior Biodegrad, 105, 233-238, doi:10.1016/j.ibiod.2015.09.013.

3. Bengelsdorf, F.R., Beck, M.H., Erz, C., Hoffmeister, S., Karl, M.M., Riegler, P., Wirth, S., Poehlein, A., Weuster-Botz, D., Dürre, P. 2018. Chapter Four - Bacterial Anaerobic Synthesis Gas (Syngas) and CO2+H2 Fermentation. in: Advances in Applied Microbiology, (Eds.) S. Sariaslani, G.M. Gadd, Vol. 103, Academic Press, pp. 143-221.

4. Chandolias, K., Pekgenc, E., Taherzadeh, M.J. 2019. Floating Membrane Bioreactors with High Gas Hold-Up for Syngas-to-Biomethane Conversion. Energies, 12(6), 1046, doi:10.3390/en12061046.

5. Christensen, N., Batstone, D.J., He, Z., Angelidaki, I., Schmidt, J.E. 2004. Removal of polycyclic aromatic hydrocarbons (PAHs) from sewage sludge by anaerobic degradation. Water Sci Technol, 50(9), 237-44, doi:10.2166/wst.2004.0580.

6. Consonni, S., Giugliano, M., Grosso, M. 2005. Alternative strategies for energy recovery from municipal solid waste Part A: Mass and energy balances. Waste Manage., 25(2), 123-35, doi:10.1016/j.wasman.2004.09.007.

21

7. Cunningham, J.A., Rahme, H., Hopkins, G.D., Lebron, C., Reinhard, M. 2001. Enhanced in situ bioremediation of BTEX-contaminated groundwater by combined injection of nitrate and sulfate. Environ Sci Technol, 35(8), 1663-70, doi:10.1021/es001722t.

8. Datar, R.P., Shenkman, R.M., Cateni, B.G., Huhnke, R.L., Lewis, R.S. 2004. Fermentation of biomass-generated producer gas to ethanol. Biotechnol Bioeng, 86, 587 - 594, doi:10.1002/bit.20071.

9. Dickhut, R.M., Andren, A.W., Armstrong, D.E. 1989. Naphthalene solubility in selected organic solvent/water mixtures. Journal of Chemical & Engineering Data, 34(4), 438-443, doi:10.1021/je00058a020.

10. Dou, J., Liu, X., Ding, A. 2009. Anaerobic degradation of naphthalene by the mixed bacteria under nitrate reducing conditions. J Hazard Mater, 165(1), 325-331, doi:10.1016/j.jhazmat.2008.10.002.

11. Drzyzga, O., Revelles, O., Durante-Rodriguez, G., Diaz, E., Garcia, J.L., Prieto, A. 2015. New challenges for syngas fermentation: towards production of biopolymers. J Chem Technol Biotechnol, 90(10), 1735-1751, doi:10.1002/jctb.4721.

12. Du, C., Xu, R., Han, S., Xu, J., Meng, L., Wang, J., Zhao, H. 2016. Solubility determination and correlation for 1,8-dinitronaphthalene in (acetone+methanol), (toluene+methanol) and (acetonitrile+methanol) mixed solvents. J Chem Thermodyn, 94, 24-30, doi:10.1016/j.jct.2015.10.015.

13. Edwards, E.A., Grbic-Galic, D. 1992. Complete mineralization of benzene by aquifer microorganisms under strictly anaerobic conditions. Appl Environ Microbiol, 58(8), 2663-6.

14. Edwards, E.A., Grbić-Galić, D. 1994. Anaerobic degradation of toluene and o-xylene by a methanogenic consortium. Appl Environ Microbiol, 60(1), 313-322.

15. Foght, J. 2008. Anaerobic Biodegradation of Aromatic Hydrocarbons: Pathways and Prospects. J Mol Microb Biotech, 15(2-3), 93-120, doi:10.1159/000121324.

16. Gershman, B. 2013. Gasification of Non-Recycled Plastics From Municipal Solid Waste In the United States. GBB Solid Waste Management Consultants.

17. Govind, R., Flaherty, P.A., Dobbs, R.A. 1991. Fate and effects of semivolatile organic pollutants during anaerobic digestion of sludge. Water Res, 25(5), 547-556, doi:10.1016/0043-1354(91)90127-C.

18. Grimalt-Alemany, A., Skiadas, I.V., Gavala, H.N. 2017. Syngas biomethanation: state-of-the-art review and perspectives. Biofuel Bioprod Bior, 12(1), 139-158, doi:10.1002/bbb.1826.

19. Heider, J., Spormann, A.M., Beller, H.R., Widdel, F. 1998. Anaerobic bacterial metabolism of hydrocarbons. FEMS Microbiol Rev, 22(5), 459-473, doi:10.1111/j.1574-6976.1998.tb00381.x.

20. Ince, O., Kolukirik, M., Cetecioglu, Z., Eyice, O., Inceoglu, O., Ince, B. 2009. Toluene inhibition on an anaerobic reactor sludge in terms of potential activity and composition of acetoclastic methanogens. J Erviron Sci Heal A, 44(14), 1551-1556, doi:10.1080/10934520903263470.

21. Liang, L., Ramesh, G. 2012. Biodegradation of Naphthalene, Benzene, Toluene, Ethyl Benzene, and Xylene in Batch and Membrane Bioreactors. Environ Eng Sci, 29(1), 42-51, doi:10.1089/ees.2010.0362.

Page 80: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

22

22. Lopes, E.J., Queiroz, N., Yamamoto, C.I., da Costa Neto, P.R. 2018. Evaluating the emissions from the gasification processing of municipal solid waste followed by combustion. Waste Manage, 73, 504-510, doi:10.1016/j.wasman.2017.12.019.

23. Maillacheruvu, K.Y., Pathan, I.A. 2009. Biodegradation of naphthalene, phenanthrene, and pyrene under anaerobic conditions. J Erviron Sci Heal A, 44(13), 1315-1326, doi:10.1080/10934520903212956.

24. Marozava, S., Mouttaki, H., Müller, H., Laban, N.A., Probst, A.J., Meckenstock, R.U. 2018. Anaerobic degradation of 1-methylnaphthalene by a member of the Thermoanaerobacteraceae contained in an iron-reducing enrichment culture. Biodegradation, 29(1), 23-39, doi:10.1007/s10532-017-9811-z.

25. Osuna, M., B., , Zandvoort, M., H.,, Iza, J., M., , Lettinga, G., Lens, P., N., L. 2003. Effects of Trace Element Addition on Volatile Fatty Acid Conversions in Anaerobic Granular Sludge Reactors. Environ Technol, 24(5), 573 - 587, doi:10.1080/09593330309385592.

26. Park, W., Jeon, C.O., Cadillo, H., DeRito, C., Madsen, E.L. 2004. Survival of naphthalene-degrading Pseudomonas putida NCIB 9816-4 in naphthalene-amended soils: toxicity of naphthalene and its metabolites. Appl Microbiol Biotechnol, 64(3), 429-35, doi:10.1007/s00253-003-1420-6.

27. Patinvoh, R.J., Osadolor, O.A., Chandolias, K., Sárvári Horváth, I., Taherzadeh, M.J. 2017. Innovative pretreatment strategies for biogas production. Bioresour Technol, 224, 13-24, doi:10.1016/j.biortech.2016.11.083.

28. Pumphrey, G.M., Madsen, E.L. 2007. Naphthalene metabolism and growth inhibition by naphthalene in Polaromonas naphthalenivorans strain CJ2. Microbiology, 153(Pt 11), 3730-8, doi:10.1099/mic.0.2007/010728-0.

29. Pöschl, M., Ward, S., Owende, P. 2010. Evaluation of energy efficiency of various biogas production and utilization pathways. Appl Energ, 87(11), 3305-3321, doi:10.1016/j.apenergy.2010.05.011.

30. Rockne, K.J., Strand, S.E. 2001. Anaerobic biodegradation of naphthalene, phenanthrene, and biphenyl by a denitrifying enrichment culture. Water Res, 35(1), 291-9, doi:10.1016/S0043-1354(00)00246-3.

31. Sahota, S., Shah, G., Ghosh, P., Kapoor, R., Sengupta, S., Singh, P., Vijay, V., Sahay, A., Vijay, V.K., Thakur, I.S. 2018. Review of trends in biogas upgradation technologies and future perspectives. Bioresour Technol Re, 1, 79-88, doi:10.1016/j.biteb.2018.01.002.

32. Shimp, R.J., Pfaender, F.K. 1985. Influence of easily degradable naturally occurring carbon substrates on biodegradation of monosubstituted phenols by aquatic bacteria. Appl Environ Microbiol, 49(2), 394-401.

33. Spath, P.L., Dayton, D.C. 2003. Preliminary Screening -Technical and Economic Assessment of Synthesis Gas to Fuels and Chemicals with Emphasis on the Potential for Biomass-Derived Syngas. National Renewable Energy Laboratory, CO, USA, NREL.

34. Stevens, D.J. 2001. Hot gas conditioning: Recent progress with larger-scale biomass gasification systems; update and summary of recent progress. National Renewable Energy Lab., Golden, CO (US).

23

35. Warthmann, R.J., Meckenstock, R.U., Schäfer, W. 2004. Inhibition of anaerobic microbial o-xylene degradation by toluene in sulfidogenic sediment columns and pure cultures. FEMS Microbiol Ecol, 47(3), 381-386, doi:10.1016/s0168-6496(03)00303-9.

36. Weelink, S.A.B., van Eekert, M.H.A., Stams, A.J.M. 2010. Degradation of BTEX by anaerobic bacteria: physiology and application. Rev Environ Sci Bio, 9(4), 359-385, doi:10.1007/s11157-010-9219-2.

37. Weiland, P. 2010. Biogas production: Current state and perspectives. Appl Microbiol Biotechnol, 85(4), 849-860, doi:10.1007/s00253-009-2246-7.

38. Westman Youngsukkasem, S., Chandolias, K., Taherzadeh, J.M. 2016. Syngas Biomethanation in a Semi-Continuous Reverse Membrane Bioreactor (RMBR). Fermentation, 2(8), 1-12, doi:10.3390/fermentation2020008.

39. Wikandari, R., Youngsukkasem, S., Millati, R., Taherzadeh, M.J. 2014. Performance of semi-continuous membrane bioreactor in biogas production from toxic feedstock containing D-Limonene. Bioresour Technol, 170, 350-5, doi:10.1016/j.biortech.2014.07.102.

40. Woolcock, P.J., Brown, R.C. 2013. A review of cleaning technologies for biomass-derived syngas. Biomass Bioenerg, 52, 54-84, doi:10.1016/j.biombioe.2013.02.036.

41. Youngsukkasem, S., Chandolias, K., Taherzadeh, M., J. 2015. Rapid bio-methanation of syngas in a reverse membrane bioreactor: Membrane encased microorganisms. Bioresour Technol, 178, 334-340, doi:10.1016/j.biortech.2014.07.071.

42. Youngsukkasem, S., Rakshit, S.K., Taherzadeh, M.J. 2012. Biogas production by encapsulated methane-producing bacteria. Bioresources, 7(1), 56-65.

Page 81: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

24

Figure captions

Fig. 1. Accumulative CH4 production, mmol; at (A) toluene concentrations of 0.5 g/L and 1.0 g/L and at (B) naphthalene concentrations of 0.1 g/L and 0.2 g/L; in RMBR and FCBR, during batch anaerobic digestion

Fig. 2 Effect of toluene on the (A) CH4 production rate, mmol CH4/(Ld); (B) CH4 production activity, (mol/molcontrol)100%; and (C) the theoretical toluene concentration, g/L; in FCBR and RMBR, during continuous experiment

Fig. 3 Effect of naphthalene on the (A) CH4 production rate, mmol CH4/(Ld); (B) CH4 production activity, (mol/molcontrol)100%; and (C) the theoretical naphthalene concentration, g/L; in FCBR and RMBR, during continuous anaerobic digestion

Fig. 4 Effect of toluene and naphthalene on the (A) CH4 production rate, mmol CH4/(Ld); (B) CH4 production activity, (mol/molcontrol)100%; and (C) the theoretical toluene and naphthalene concentration, g/L; in FCBR and RMBR, during continuous anaerobic digestion

25

Fig. 1

Page 82: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

26

Fig. 2

27

Fig. 3

Page 83: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

28

Fig. 4

78

78

Paper V

Floating Membrane Bioreactors with High Gas Hold-Up for Syngas-to-Biomethane

Conversion

Page 84: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

energies

Article

Floating Membrane Bioreactors with High GasHold-Up for Syngas-to-Biomethane Conversion

Konstantinos Chandolias 1,* , Enise Pekgenc 2 and Mohammad J. Taherzadeh 1

1 Swedish Centre for Resource Recovery (SCRR), Faculty of Textiles engineering and Business, University ofBorås, Allégatan 1, 501 90 Borås, Sweden; [email protected]

2 Department of Environmental Engineering, Istanbul Technical University, Maslak,34467 Sarıyer/Istanbul, Turkey; [email protected]

* Correspondence: [email protected]

Received: 20 February 2019; Accepted: 14 March 2019; Published: 18 March 2019����������������

Abstract: The low gas-to-liquid mass transfer rate is one of the main challenges in syngasbiomethanation. In this work, a new concept of the floating membrane system with high gas hold-upwas introduced in order to enhance the mass transfer rate of the process. In addition, the effectof the inoculum-to-syngas ratio was investigated. The experiments were conducted at 55 ◦C withan anaerobic mixed culture in both batch and continuous modes. According to the results fromthe continuous experiments, the H2 and CO conversion rates in the floating membrane bioreactorwere approximately 38% and 28% higher in comparison to the free (suspended) cell bioreactors.The doubling of the thickness of the membrane bed resulted in an increase of the conversion rates ofH2 and CO by approximately 6% and 12%, respectively. The highest H2 and CO consumption ratesand CH4 production rate recorded were approximately 22 mmol/(L·d), 50 mmol/(L·d), and 34.41mmol/(L·d), respectively, obtained at the highest inoculum-to-syngas ratio of 0.2 g/mL. To conclude,the use of the floating membrane system enhanced the syngas biomethanation rates, while a thickermembrane bed resulted in even higher syngas conversion rates. Moreover, the increase of theinoculum-to-syngas ratio of up to 0.2 g/mL favored the syngas conversion.

Keywords: floating MBR; syngas-to-biomethane conversion; high gas hold-up; inoculum-to-syngas ratio

1. Introduction

Gasification is a thermochemical process which converts biomass into a gaseous mixture,called syngas (mostly CO, H2, and CO2). This gas can be employed for the production of electricity,energy, and transport fuels. Currently, approximately 50% of the generated syngas is converted intoNH3, 25% into H2, and the remaining fraction is used for the production of Fischer Tropsch fuels,methanol, and other valuable chemicals [1,2]. Syngas is also considered as a promising vector for heatand power generation [3]. Another promising application of syngas and other industrial off-gases isanaerobic fermentation and conversion into biofuels, alcohols, bioplastics, and value-added chemicals.Syngas biomethanation, in particular, is considered as a sustainable alternative to the applicationsmentioned above [4].

An important limitation of syngas fermentation is the low syngas conversion rate due to poorgas-to-liquid mass transfer rates. Different approaches, in order to improve the mass transfer,focus mainly on the bioreactor’s design and the syngas feed. More specifically, different bioreactortypes such as stirred tank, bubble, packed bed, airlift, and trickle bed bioreactors have been studied.Packed bed and membrane bioreactors allowed higher mass transfer and cell-density than thetraditional bubble column and stirred tank reactors during syngas fermentation [5]. In addition,

Energies 2019, 12, 1046; doi:10.3390/en12061046 www.mdpi.com/journal/energies

Page 85: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

Energies 2019, 12, 1046 2 of 14

the use of microbubble sparging reduced the power-to-energy demand [6]. Hollow fibre membraneswere also employed for the syngas feed and led to high mass transfer efficiency [7]. Another effect ofthe hollow fibre membranes was the creation of a biofilm on the surface of the membranes for bettergas-cell contact [8].

Another way to increase the gas-to-liquid mass transfer without using costly andenergy-demanding methods, such as agitation, is to increase the gas hold-up. Higher gas hold-upallows for better diffusion of the gas inside the bulk medium and better contact with the cells. This canbe achieved by altering the bioreactor´s design [4], such as by increasing the height and decreasingthe diameter of the bioreactor or by blocking the anodic path of the rising gas bubbles inside theliquid medium.

The main novelty in this work is in the function of polymeric membranes which were employedin order to decelerate the anodic bubble-rise velocity and thus increase the gas hold-up of syngas in afloating membrane bioreactor (floating MBR). Initially, the membranes were shaped into rectangularsachets which were seeded with anaerobic cells and then heat-sealed and placed inside the medium ofthe floating MBR. The membrane-floating effect was caused by the biogas that was produced insidethe membranes. The swollen membranes formed a packed membrane bed which floated under thesurface of the liquid medium with the help of a plastic net. In other words, the floating MBR systemis actually a reverse membrane bioreactor (reverse MBR) [9] in which the membranes form a packedfloating bed inside the medium of the reactor.

Another factor that was studied in this work and can affect the syngas conversion rate is theinoculum-to-syngas ratio (ISR). Although the effect of ISR on the methane potential of organic wastehas been investigated in other studies [10,11], there is no previous study on the effect of the ISR insyngas biomethanation. This parameter could be essential for the improvement of the process asa low ratio could lead to toxic syngas concentrations for the cells, while a high ratio could causecell-starvation. For this purpose, several ISRs were tested in both batch and continuous experiments.

The aim of this work was to investigate the effect of a MBR with a floating membrane bed on thesyngas conversion rate and the effect of increasing ISR during syngas biomethanation. According tothe authors’ best knowledge, there are no similar studies in the literature.

2. Materials and Methods

2.1. Inoculum, Syngas, and Liquid Medium

The inoculum was a mixed culture collected from a local thermophilic 3000 m3 anaerobic digesteroperating on organic fraction of food solid waste (Borås Energy and Environment AB, Borås, Sweden).Prior to the experiment, the inoculum was incubated for four days at 55 ◦C in order to consume all thenutrients. The pH of the inoculum was 8.0 and the total solid (% TS) and volatile solid (% VS) contentwas 15.01 ± 0.17% and 65.11 ± 0.44%, respectively.

The gaseous substrate was syngas, containing 20% H2, 55% CO, and 10% CO2, and was providedby AGA gas AB (Gothenburg, Sweden). The gas cylinder, containing the syngas, was pressurized withnitrogen gas at 1500 psi.

The liquid medium consisted of basal medium (micronutrients and macronutrients) and aceticacid. The basal medium recipe was obtained from the literature [12]. The macronutrients concentrationin the liquid medium was 280 mg NH4Cl/L, 330 mg K2HPO4·3H2O/L, 100 mg MgSO4·7H2O/L,and 10 mg CaCl2·2H2O/L. In addition, 2 mL of trace elements solution was added per 1 L of liquidmedium. The composition of the trace elements solution was 2000 mg FeCl2·4H2O/L, 50 mg H3BO3/L,50 mg ZnCl2/L, 500 mg MnCl2·4H2O/L, 38 mg CuCl2·2H2O/L, 50 mg (NH4)6Mo7O24·4H2O/L, 2000mg CoCl2·6H2O/L, 142 mg NiCl2·6H2O/L, and 164 mg Na2SeO3·5H2O/L.

Energies 2019, 12, 1046 3 of 14

2.2. Bioreactor Characteristics, Seeding, and Start Up

The excess water of the inoculum was removed by centrifugation at 9000× g for 3 min (HereausMegafuge 8, Thermo Scientific, Bremen, Germany) prior to seeding. Both batch and continuousexperiments were conducted. The batch bioreactors were serum glass bottles with a rubber sealed capand a total volume of 118 mL (Bioprocess control AB, Lund, Sweden). Each bioreactor was loadedwith 40 mL of liquid medium and different amounts of inoculum in order to investigate the effect ofdifferent ISRs of 0.01, 0.04, 0.07, and 0.1 g/mL. After seeding, the bioreactors were placed inside ashaking water bath at 100 rpm, at an inclination of approximately 45◦, and a temperature of 55 ◦C.

The bioreactors that were operated in the continuous mode were glass bubble column vesselswith rubber sealing caps and a total volume of 2.2 L. These bioreactors were placed inside a stationarywater bath at 55 ◦C. Each bubble column reactor contained 1.7 L of liquid medium and was seededwith 45 g of the pelleted inoculum. The continuous experiments were conducted in order to comparethe floating MBR with other types of free cell bioreactors and the effect of thicker membrane bed andhigher ISR. More specifically, the bioreactor designs that were used were a floating MBR; a bioreactorwith both floating membrane bed and free cells (floating MBR/FCBR); a bioreactor with free cells,filled with packing material (PBR); and a free cell bioreactor (FCBR). The FCBR had an ISR of 0.1 g/mLand was operated in parallel with another free cell bioreactor (FCBR.2), which was loaded with 90 ginoculum (ISR = 0.2 g/mL). The Floating MBR contained 15 membranes filled with 3 g of inoculumeach. In the floating MBR/FCBR, a part of the inoculum (24 g) was encased inside eight membranesand another part (26 g) was suspended in the liquid medium of the reactor. The PBR and FCBRcontained free cells with the difference that the PBR was filled with 1.5 L of packed material, designedfor cell-attach growth, with a diameter of 15 mm, made from PVC (HR 15–7, Rauschert, Hannover,Germany). Finally, another bioreactor, the floating MBR.2, was loaded with 30 membrane sachets (1.5g inoculum/membrane), forming a floating membrane bed with a thickness of approximately 6 cm,while the thickness of the membrane bed in the floating MBR was approximately 3 cm.

The membranes were hydrophilic polyvinylidene difluoride (PVDF) flat sheets (Merck MilliporeLtd., Cork, Ireland) with a pore size and thickness of 0.1 μm and 125 μm, respectively, while themethod of cell-encasement was described in a previous work [13].

2.3. Gas and Liquid Feeding

Figure 1 shows a scheme of the gas and liquid feeding process. Syngas was fed in the bioreactorswith a flow meter control (11–110 mL/min, Swagelok, Sollentuna, Sweden) at a flow rate of 15 mL/min.The gaseous substrate was introduced to the bottom of the bioreactors with a sparger (Air diffuser2 × 3 cm, Zalux, Zaragoza, Spain). There was no gas-recirculation and the conversion rate of syngaswas reported after one pass through the medium broth. The gas composition was analysed at theinlet and outlet of the reactors, while the gas flow rate at the outlet was recorded by a data acquisitionsystem (AMPTS II, Bioprocess control, Sweden AB, Sweden). The feeding of the liquid medium tookplace with a plastic 50 mL syringe though a tube that was immersed inside the medium. Regularcontrols with gas and liquid sampling ensured the stable operation of the bioreactors.

Page 86: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

Energies 2019, 12, 1046 4 of 14

Figure 1. Schematic of gas and liquid feeding in the floating MBR (membrane bioreactor).

2.4. Analytical Methods

The gas components were detected with Gas Chromatography. The CO levels were obtained bya Gas Chromatograph (Perkin-Elmer 480, Norwalk, CT, USA) with a packed column (CarboxenTM1000, SUPELCO, 6� × 1.8” OD, 60/80 Mesh, Shelton, CT, USA) and a thermal conductivity detector(Perkin-Elmer) with an injection temperature of 200 ◦C. The H2, CO2, and CH4 levels were analysedwith a Gas Chromatograph (Perkin-Elmer 590, Norwalk, CT, USA), equipped with a packed column(CarboxenTM 1000, SUPELCO, 6� × 1.8” OD, 60/80 Mesh, Shelton, CT, USA), using a thermalconductivity detector (Perkin-Elmer, Norwalk, CT, USA) with an injection temperature of 200 ◦C.The gas sampling was performed with a 0.25 mL gas-tight syringe (VICI, Precision Sampling Inc., BatonRouge, LA, USA) and the carrier gas in the chromatographs was N2 with a flow rate of 30 mL/min at75 ◦C.

For the analysis of the liquid samples, a High-Performance Liquid Chromatography (Waters 2695,Waters Corporation, Milford, CT, USA) with a hydrogen-based column (Aminex HPX87-H, BioRadLaboratories, München, Germany), equipped with a refractive index (RI) detector (Waters 2410, WatersCorporation, Milford, CT, USA), was operated at 60 ◦C and 0.6 mL/min (5 mM H2SO4 eluent).

The gas analysis of the continuous experiment took place every five days starting from the 50thday and ending on the 90th day of fermentation, while the batch experiment was run in triplicatesfor a period of seven days and gas samples were analysed every 24 h. The results are presented asaverage value ± standard deviation.

3. Results and Discussion

A main challenge during syngas fermentation is the low gas-to-liquid mass transfer, which can beimproved by increasing the gas hold-up inside the liquid medium of bioreactors. For this purpose,anaerobic cells were encased in membrane sachets which were then heat-sealed and immersed insidea bubble column bioreactor. During the fermentation process, syngas was introduced to the bottomof the bioreactor with a sparger and the bubble ascent of syngas components was delayed because

Energies 2019, 12, 1046 5 of 14

of the membrane bed, which operated as a mechanical barrier. This delay increased the gas hold-upin the liquid phase. Consequently, the syngas components diffused in the liquid phase and throughthe 0.1 μm membrane pores. The total interfacial area for gas mass transfer (α) increases with highergas hold-up and smaller bubble size. Thus, higher mass transfer rates are achieved in systems withhigher gas hold-up and smaller bubble size [14]. In addition, the position of the membrane bed underthe liquid surface and the gas passing through the membrane pores and into the cell-contained areafavoured the gas-to-cell contact without agitation.

The membranes were hydrophilic and, therefore, accessible to dissolved syngas componentsand organic acids in the liquid phase. These components diffused through the membrane pores andthereafter were converted into biogas. The produced biogas built a pressure inside the membraneswhich increased until it reached the membrane bubble point. Then, biogas exited the inner membranearea by blowing the membrane pores dry. Consequently, the inner pressure of the membrane droppedagain, and fresh liquid diffused through the membrane. The fact that the membrane sachets werecontinuously swollen led also to the conclusion that there was probably minimum gas exchangefrom the outside to the inside of the sachets. The membranes are accessible to gas when they aredry, however once they are wet, the differential pressure has to be greater than the bubble point for agas exchange to occur. This means that the membranes worked as conductors of the diffused liquidphase and the produced gas phase. From previous studies, it was observed that the floating effect wasdirectly stirred by the organic loading rate (OLR) of the bioreactors [15] because the biogas productionrate inside the membranes had to be high in order to create the floating effect. The experiments inthis work showed that a minimum OLR of approximately 1 g COD/(L·d) was required in order toinitiate the membrane floating phenomenon. The undissolved syngas bubbles ascended between themembrane sachets to the surface of the liquid.

Figure 2 illustrates the different bioreactor designs that were investigated. The floating MBRwas operated in parallel with a free cell bioreactor (FCBR) containing free suspended cells, a packedbioreactor (PBR) containing free cells growing on packed material, and a hybrid bioreactor (floatingMBR/FCBR) containing both membrane encased cells and free cells. These bioreactors are describedin more detail in the Materials and Methods section. The effect of a thicker membrane bed was alsoinvestigated as well as the effect of a higher inoculum-to-syngas ratio (ISR). The H2 and CO conversionrates and the CO2 and CH4 production rates in mmol/(L·d) were used as indicators of the reactors’efficacy. The results showed that the use of the membrane bed resulted in higher syngas conversionrates. The increase of the thickness of the membrane bed led to higher conversion rates. In addition,higher ISRs increased the conversion rates in both batch and continuous experiments.

During the starting up period (first 30 days) of the continuous experiment, the bioreactors werefed with basal medium which contained micro- and macro-nutrients and syngas as the sole carbonand energy source. From the 31st day and until the 90th day, acetic acid was also introduced into theliquid medium with an OLR of 1.21 g COD/(L·d). The HRT was 34 d during the experiment.

Page 87: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

Energies 2019, 12, 1046 6 of 14

Figure 2. Syngas biomethanation in bioreactors containing (a) membrane-encased cells (floating MBR),membrane-encased and free cells (floating MBR/FCBR), free cells growing on packed material (PBR),and free (suspended) cells (FCBR); (b) membrane bed thickness of 3 cm (floating MBR) and 6 cm(floating MBR.2), and (c) inoculum to syngas ratio (ISR) of 0.1 g/mL (FCBR) and 0.2 g/mL (FCBR.2).

3.1. Efficacy of the Membrane Floating Bed System

The inoculum used in this work consisted of mixed cells which take up syngas from the liquidphase. Therefore, the rate of the mass transfer and, thus, the mass transfer coefficient from thegas-to-liquid phase is of vital importance. The syngas components face several resistances during theirjourney from the gas phase to the cells (Figure 3a). In the case of floating MBR, there is an extra masstransfer resistance of the membrane surface, however previous studies proved that this resistance isnegligible [15]. The liquid film surrounding the cells (Figure 3b) is considered to be the main resistanceof the gas-to-liquid mass transfer inside the investigated bioreactors, while the rest of the resistancesare considered negligible [14]. The following equations show the rate of mass transfer of component Athough the gas (NAG) and the liquid (NAL) boundary in gmol/(m3·s):

NAG = kGα(CAG − CAGi), (1)

NAL = kLα(CALi − CAL), (2)

where kG and kL is the gas-phase and liquid-phase mass transfer coefficient, m/s, respectively, and α

is the total interfacial area for mass transfer, 1/m. The concentration of A in the liquid bulk is CAL and

Energies 2019, 12, 1046 7 of 14

in the liquid boundary is CAli, while the concentration of A in the gas phase is CAG and in the gasboundary is CAGi, in gmol/m3.

Figure 3. (a) The resistances during the mass transfer of syngas from the gas phase to the site ofreaction in the cells. 1. Travel in the gas bubble; 2. Move across the gas-liquid interfacial; 3. Travelthrough the liquid film surrounding the gas bubble; 4. Travel in the liquid bulk; 5. Enter the interiorof the membrane through its pores; 6. Move inside the membrane; 7. Travel across the liquid filmsurrounding the microbial cell; 8. Pass through the cell membrane; 9. Move through the cell and endup in the site of reaction; (b) Movement of A through the interfacial boundary. CAL: concentration of Ain the liquid phase; CAli concentration of A in the liquid boundary; CAGi concentration of A in the gasboundary; and CAG concentration of A in the gas phase [14].

Equations (1) and (2) can be simplified to Equations (3) and (4) [14], respectively, where mCAL

= C*AG is the gas-phase concentration of A in equilibrium with CAL and CAG/m = C*

AL is theliquid-phase concentration of A in equilibrium with CAG, and m is the distribution factor.

NAG = kGα(CAG − C*AG), (3)

NAL = kLα(C*AL − CAL), (4)

Equations (3) and (4) are valid for systems in which the main mass transfer resistance is either thegas-phase film resistance or the liquid-phase film resistance. Thus, the overall mass transfer coefficientsKGα and KLα were replaced by kGα and kLα [14]. In the case of syngas fermentation, H2 and COare poorly soluble to the aqueous solution, which means that kGα is significantly larger than kLα

and, therefore, Equation (3) is the main equation that can describe the limiting mass transfer rate inthe system.

The consumption rates of H2 and CO and the production rates of CH4 and CO2 in floating MBR,floating MBR/FCBR, PBR, and FCBR are presented in Figure 4. The highest consumption rates ofH2 and CO obtained were 18.47 ± 1.25 and 39.67 ± 1.45 mmol/(L·d), respectively, in floating MBR.In the same bioreactor, the highest CH4 production of 21.33 mmol CH4/(L·d) was also achieved

Page 88: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

Energies 2019, 12, 1046 8 of 14

and the highest CH4 yield of 0.36 mol CH4/mol (H2 + CO). The pH of the effluent was 8.2 ± 2.0in all bioreactors during the experiment. In previous studies, a triculture (R. rubrum, M. barkeri, M.formicicum) converted syngas of a similar composition into CH4 in a trickle bed bioreactor with a CH4

production rate of 48–72 mmol CH4/(L·d) and yields of 0.2–0.214 mol CH4/mol (H2 + CO) [16,17].Moreover, the same triculture converted syngas in a packed bed bioreactor with a CH4 productionof 4.8–7.2 mmol/(L·d) and a yield of 0.214 mol CH4/mol (H2 + CO) [17]. Higher yields of 0.6–0.8mol CH4/mol (H2 + CO) and a CH4 production of 73 mmol/(L·d) were obtained in a multi-orificeoscillatory baffled bioreactor where granular sludge converted syngas with a gas recirculation rate of600 mL/min [18].

Figure 4. Syngas biomethanation in floating MBR, floating MBR/FCBR, PBR, and FCBR duringcontinuous syngas feeding. Consumption of (a) H2 and (b) CO and production of (c) CO2 and(d) CH4in mmol/(L·d).

The lowest syngas biomethanation rates were observed in the PBR and the FCBR. This resultcontradicts with other studies, which reported that the use of packed material in the PBR significantlyimproved the syngas conversion rates and CH4 production rate. For instance, Burkhardt, et al. [19]reported a CH4 concentration of higher than 98% in their final biogas product, achieved in a trickle bedbioreactor filled with packing material. Trickle bed bioreactors with packed material are consideredeffective in syngas biomethanation because of higher mass transfer coefficients and lower cost thanthe continuous stirred bioreactors (CSTR) [16]. The liquid recirculation, which may be co-current orcounter-current with the gas flow, is one of the main reasons for high mass transfer in the trickle bedbioreactors. However, in the present study, there was no liquid recirculation and, therefore, the lowersyngas conversion rates in PBR are likely a result of inadequate substrate distribution inside thebioreactor, which is probably the reason for the lower acetic acid consumption (Table 1) in comparisonto other bioreactors.

Energies 2019, 12, 1046 9 of 14

Table 1. Consumption of acetic acid in the bioreactors.

Bioreactor Type Acetic Acid Consumed, g/(L·d)

floating MBR (membr. bed = 3 cm) 1.40 ± 0.42floating MBR/FCBR 1.01 ± 0.41

PBR 0.78 ± 0.51FCBR (ISR = 0.1 g/mL) 0.92 ± 0.52

floating MBR.2 (membr. bed = 6 cm) 1.15 ± 0.35FCBR.2 (ISR = 0.2 g/mL) 1.79 ± 0.62

The floating MBR/FCBR system proved to be less efficient in terms of syngas conversion incomparison to the floating MBR system. The reason for this was probably the better gas-to-cell contactin floating MBR due to the fact that the suspended inoculum was mostly concentrated at the bottom ofthe floating MBR/FCBR. Likewise, in the FCBR and the PBR, the inoculum was mostly concentratedat the bottom of the bioreactors. In bubble column bioreactors with free cells, when syngas feed isintroduced at the bottom of the bioreactor, the H2 and CO levels decrease as the gas flows up thecolumn because of cellular consumption. This causes spatial difference of dissolved syngas inside thebubble column which can affect the cellular growth and the product profile [20]. Therefore, the aim isto establish favourable syngas concentration profiles in the liquid medium according to the desirableproduct [20]. However, in the floating MBR, the cells were assembled in packed formation at the upperparts of the liquid medium so that the gas bubbles stayed longer in the liquid and were dispersedthough the membrane pores. This cell-placement was probably a main factor which caused bettergas-to-cell contact and, therefore, faster syngas biomethanation in the floating MBR.

The floating MBR resembles the concept of a trickle bed or a packed bed bioreactor where thepacked material is replaced by membranes. Trickle bed bioreactors are preferred to other bioreactortypes, such as continuous stirred bioreactors, because of their enhanced gas-to-liquid mass transfermechanisms. A mass transfer analysis on various bioreactors claimed that the mass transfer wasenhanced over three times in a trickle bed bioreactor in comparison to continuously stirred tankreactor [21]. In addition, the optimum CO mass transfer coefficient of stirred tank bioreactors can besubstantially improved in bubble column bioreactors because of higher mass transfer driving forces,which are a result of gas composition spatial profiles and longer gas hold-up [20].

3.2. The Effect of Membrane Bed Thickness

The floating MBR, containing a floating membrane bed, presented the highest syngas conversionefficacy in comparison to other bioreactor types in the previous section. In order to study the effectof membrane bed thickness, another bioreactor was operated in parallel with the floating MBR. Thissecond bioreactor (floating MBR.2) consisted of a higher amount of membranes (30) that was twotimes the amount of membranes in the floating MBR. The floating membrane bed in floating MBR.2was approximately 6 cm thick in comparison with the membrane bed in floating MBR which had athickness of approximately 3 cm. The results in Figure 5 show that the reinforcement of the membranebed increases the conversion rates of H2 and CO by approximately 6% and 11%, respectively, in floatingMBR.2. This result proved that bioreactors with a thicker membrane bed had higher conversion rates,although the improvement was not proportional to the magnitude of the increase of the thickness ofthe membrane bed, which was approximately 100%.

Page 89: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

Energies 2019, 12, 1046 10 of 14

Figure 5. Syngas biomethanation in floating MBR and floating MBR.2 with a membrane bed thicknessof 3 cm and 6 cm, respectively, during continuous syngas feeding. Consumption of (a) H2 and (b) COand production of (c) CO2 and (d) CH4 in mmol/(L·d).

3.3. Impact of Inoculum-to-Syngas Ratio (ISR)

In order to study the effect of different ISRs, both batch and continuous bioreactors were operated.During the batch experiment, syngas consumption and biogas production was investigated at differentISRs (0.01, 0.04, 0.07, and 0.1 g/mL), and the results are presented in Figure 6. According to thefigure, the increase of ISR led to faster consumption of H2 and CO. More specifically, the completeconsumption of H2 and CO took one day in bioreactors with an ISR of 0.1 g/mL, while in bioreactorswith 10 times less inoculum (ISR = 0.01 g/mL), the complete consumption of H2 and CO took fourand five days, respectively. The CH4 production followed the same trend. The CH4 production inthe bioreactor with an ISR of 0.1 g/mL was stabilized on the third day, while the CH4 productionin the bioreactor with an ISR of 0.01 g/mL reached its highest CH4 production on the sixth day offermentation. In bioreactors with an ISR of 0.04 g/mL, the complete conversion of H2 and CO tookthree and five days, respectively. These conversion rates were lower in comparison with a previousstudy with a similar system (ISR = 0.38 g/mL) during which H2 and CO were totally consumed intwo days [22]. However, in this work, the initial amount of H2 and CO content in the bioreactors wasapproximately 50% and 24% higher than in the previous study. The highest CH4 yield (at an ISR of0.1 g/mL) obtained in the current work was 0.39 mol CH4/mol (H2 + CO2), which is comparativelyhigher than in other similar studies. For example, the CH4 yield achieved during the conversion ofH2/CO2 and CO in a batch bubble column bioreactor with an ISR of 0.0037 g/mL was 0.22–0.26 molCH4/mol (H2 + CO2) and 0.25 mol CH4/mol CO, respectively [23]. The lower yields in the abovework could have been caused by the lower ISR and the different syngas composition in comparison tothe current work.

Energies 2019, 12, 1046 11 of 14

Figure 6. Syngas biomethanation in batch bioreactors with an ISR of 0.01, 0.04, 0.07, and 0.1 (g/mL).Consumption of (a) H2 and (b) CO and production of (c) CO2 and (d) CH4 in mmol.

During the continuous experiment, the FCBR that had an ISR of 0.1 g/mL was operated inparallel with FCBR.2 with an ISR of 0.2 g/mL. The aim was to investigate the effect of doubling theISR and the operation in continuous mode. The results (Figure 7) showed an increase in the H2 andCO consumption rate of approximately 66% and 61%, respectively, and an increase of approximately94% in CH4 production rate. The CH4 production rate in the bioreactor with an ISR of 0.2 g/mLwas 34.41 mmol/(L·d), the highest achieved in this work. This production rate is comparable toresults from other similar studies. A CH4 production rate of 72 mmol/(L·d) was achieved duringbiomethanation of syngas with a similar composition at a gas flow rate of 70 mL/min. The syngas wasconverted in a trickle bed bioreactor by a triculture of R. rubrum, M. barkeri, and M. formicicum [4,16].Another study was conducted in a packed bed bioreactor with the same triculture, similar syngascomposition, and gas flow rate of 80 mL/min. In that work, a lower CH4 production rate between 4.8and 7.2 mmol/(L·d) was reported [17]. In the literature, there are previous studies that have reportedthat higher ISRs can improve the CH4 potential yields. These studies have mainly focused on theinvestigation of using different ISRs based on g VSadded, with mixed anaerobic sludge. For example, astudy on the effect of ISR on the CH4 potential of microcrystalline cellulose production wastewaterreported that the fastest CH4 production rate and highest kinetic constant were achieved at the highestISR of 2.0 [10]. However, extremely high ISRs may be inhibiting for the anaerobic process. Lim andFox [24] studied three different ISRs (1, 0.33, and 0.125) and reported that the highest CH4 productionrate was obtained at the ratio of 0.33, whereas the minimum production rate was obtained at an ISRof 0.125. The low ISR probably caused low substrate concentration and the high ISR caused highconcentration of volatile fatty acids and thus, low pH [24].

Page 90: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

Energies 2019, 12, 1046 12 of 14

Figure 7. Syngas biomethanation in FCBR (ISR = 0.1 g/mL) and FCBR.2 (ISR = 0.2 g/mL) duringcontinuous syngas feeding. Consumption of (a) H2 and (b) CO and production of (c) CO2 and (d) CH4

in mmol/(L·d).

4. Conclusions

The new concept of floating MBR was successfully applied during continuous syngasbiomethanation with a mixed culture and thermophilic conditions. The floating MBR was operatedin parallel with a hybrid bioreactor containing both membrane encased and free cells, a bioreactorwith free cells growing on packed material, and a free cell bioreactor. The results showed thatthe use of the floating MBR improved the gas hold-up and accelerated the syngas conversionrate. Thus, syngas conversion rates were higher in the floating MBR in comparison to the otherbioreactors. The increase of the thickness of the membrane bed by twofold resulted in higher syngasconversion rates of 6% and 11% for H2 and CO consumption, respectively. In addition, differentinoculum-to-syngas ratios (ISR) were tested during both batch and continuous experiments and thehighest syngas conversion rates were achieved at the highest ISR of 0.2 g/mL during continuousbiomethanation. The study of the limiting effect of ISR in continuous syngas biomethanation isconsidered an interesting future step.

Author Contributions: Conceptualization, K.C.; methodology, K.C. and E.P.; validation, K.C.; formal analysis,K.C.; investigation, E.P. and K.C.; resources, E.P. and K.C.; data curation, E.P. and K.C.; writing—original draftpreparation, K.C.; writing—review and editing, K.C. and M.J.T.; visualization, K.C. and E.P.; supervision, M.J.T.;project administration, K.C. and M.J.T.; funding acquisition, M.J.T.

Funding: This research was funded by the Swedish Research Council.

Acknowledgments: The authors would like to thank Jean-Francois Charlot from Merck, Millipore S.A.S., Franceand Amir Mahboubi from University of Borås, Sweden for their technical support on membranes.

Conflicts of Interest: The authors declare no conflict of interest.

References

1. Rauch, R.; Hrbek, J.; Hofbauer, H. Biomass gasification for synthesis gas production and applications of thesyngas. Wiley Interdiscip. Rev. Energy Environ. 2014, 3, 343–362. [CrossRef]

Energies 2019, 12, 1046 13 of 14

2. Boerrigter, H.; Rauch, R.; Knoef, H.A.M. Handbook Biomass Gasification, 2nd ed.; BTG Biomass TechnologyGroup: Enschede, The Netherlands, 2005.

3. Solarte-Toro, J.C.; Chacón-Pérez, Y.; Cardona-Alzate, C.A. Evaluation of biogas and syngas as energy vectorsfor heat and power generation using lignocellulosic biomass as raw material. Electron. J. Biotechnol. 2018, 33,52–62. [CrossRef]

4. Grimalt-Alemany, A.; Skiadas, I.V.; Gavala, H.N. Syngas biomethanation: State-of-the-art review andperspectives. Biofuels Bioprod. Biorefin. 2017, 12, 139–158. [CrossRef]

5. Asimakopoulos, K.; Gavala, H.N.; Skiadas, I.V. Reactor systems for syngas fermentation processes: A review.Chem. Eng. J. 2018, 348, 732–744. [CrossRef]

6. Bredwell, M.D.; Srivastava, P.; Worden, R.M. Reactor Design Issues for Synthesis-Gas Fermentations.Biotechnol. Progr. 1999, 15, 834–844. [CrossRef] [PubMed]

7. Shen, Y.; Brown, R.; Wen, Z. Enhancing mass transfer and ethanol production in syngas fermentationof Clostridium carboxidivorans P7 through a monolithic biofilm reactor. Appl. Energy 2014, 136, 68–76.[CrossRef]

8. Shen, Y.W.; Brown, R.; Wen, Z.Y. Syngas fermentation of Clostridium carboxidivoran P7 in a hollow fibermembrane biofilm reactor: Evaluating the mass transfer coefficient and ethanol production performance.Biochem. Eng. J. 2014, 85, 21–29. [CrossRef]

9. Mahboubi, A.; Ylitervo, P.; Doyen, W.; De Wever, H.; Taherzadeh, M.J. Reverse membrane bioreactor:Introduction to a new technology for biofuel production. Biotechnol. Adv. 2016, 34, 954–975. [CrossRef][PubMed]

10. Rodriguez-Chiang, L.M.; Dahl, O.P. Effect of Inoculum to Substrate Ratio on the Methane Potential ofMicrocrystalline Cellulose Production Wastewater. Bioresources 2015, 10, 898–911. [CrossRef]

11. Pellera, F.M.; Gidarakos, E. Effect of substrate to inoculum ratio and inoculum type on the biochemicalmethane potential of solid agroindustrial waste. J. Environ. Chem. Eng. 2016, 4, 3217–3229. [CrossRef]

12. Osuna, M.B.; Zandvoort, M.H.; Iza, J.M.; Lettinga, G.; Lens, P.N.L. Effects of Trace Element Addition onVolatile Fatty Acid Conversions in Anaerobic Granular Sludge Reactors. Environ. Technol. 2003, 24, 573–587.[CrossRef] [PubMed]

13. Youngsukkasem, S.; Rakshit, S.K.; Taherzadeh, M.J. Biogas production by encapsulated methane-producingbacteria. Bioresources 2012, 7, 56–65.

14. Doran, P.M. Mass transfer. In Bioprocess Engineering Principles, 2nd ed.; Elsevier: Amsterdam,The Netherlands, 2012; pp. 379–444.

15. Westman, Y.S.; Chandolias, K.; Taherzadeh, J.M. Syngas Biomethanation in a Semi-Continuous ReverseMembrane Bioreactor (RMBR). Fermentation 2016, 2, 8. [CrossRef]

16. Kimmel, D.E.; Klasson, K.T.; Clausen, E.C.; Gaddy, J.L. Performance of trickle-bed bioreactors for convertingsynthesis gas to methane. Appl. Biochem. Biotechnol. 1991, 28–29, 457–469. [CrossRef]

17. Klasson, K.T.; Cowger, J.P.; Ko, C.W.; Vega, J.L.; Clausen, E.C.; Gaddy, J.L. Methane production from synthesisgas using a mixed culture ofR. rubrum M. barkeri, and M. formicicum. Appl. Biochem. Biotechnol. 1990, 24,317–328. [CrossRef]

18. Pereira, F.M.R. Intensified Bioprocess for the Anaerobic Conversion of Syngas to Biofuels. Ph.D. Thesis,University of Minho, Braga, Portugal, 2014. Available online: http://hdl.handle.net/1822/34834 (accessedon 18 March 2019).

19. Burkhardt, M.; Koschack, T.; Busch, G. Biocatalytic methanation of hydrogen and carbon dioxide in ananaerobic three-phase system. Bioresour. Technol. 2015, 178, 330–333. [CrossRef] [PubMed]

20. Chen, J.; Gomez, J.A.; Höffner, K.; Barton, P.I.; Henson, M.A. Metabolic modeling of synthesis gasfermentation in bubble column reactors. Biotechnol. Biofuels 2015, 8, 89. [CrossRef] [PubMed]

21. Atiyeh, H.K.; Devarapalli, M.; Lewis, R.S.; Huhnke, R.L. Semi-Continuous Syngas Fermentation in a TrickleBed Reactor. In Proceedings of the AIChE 2013 Annual Meeting, San Franscisco, CA, USA, 3–8 November2013.

22. Youngsukkasem, S.; Chandolias, K.; Taherzadeh, M.J. Rapid bio-methanation of syngas in a reversemembrane bioreactor: Membrane encased microorganisms. Bioresour. Technol. 2015, 178, 334–340. [CrossRef][PubMed]

Page 91: Enhanced Methane and Hydrogen production in Reverse ...hb.diva-portal.org/smash/get/diva2:1352281/FULLTEXT01.pdf2 production, during syngas fermentation (paper III). In addition, I

Energies 2019, 12, 1046 14 of 14

23. Bugante, E.C.; Shimomura, Y.; Tanaka, T.; Taniguchi, M.; Oi, S. Methane production from hydrogen andcarbon dioxide and monoxide in a column bioreactor of thermophilic methanogens by gas recirculation.J. Ferment Bioeng. 1989, 67, 419–421. [CrossRef]

24. Lim, S.J.; Fox, P. Biochemical methane potential (BMP) test for thickened sludge using anaerobic granularsludge at different inoculum/substrate ratios. Biotechnol. Bioprocess. Eng. 2013, 18, 306–312. [CrossRef]

© 2019 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open accessarticle distributed under the terms and conditions of the Creative Commons Attribution(CC BY) license (http://creativecommons.org/licenses/by/4.0/).