+ All Categories
Home > Documents > Directing the metabolism of xylose towards P(3HB), xylitol ... · The work described in this...

Directing the metabolism of xylose towards P(3HB), xylitol ... · The work described in this...

Date post: 21-Jan-2019
Category:
Upload: trinhnhan
View: 214 times
Download: 0 times
Share this document with a friend
108
Xylose Upgrade Directing the metabolism of xylose towards P(3HB), xylitol and xylonic acid in Burkholderia sacchari Inˆ es Maria Vicente Palolo Thesis to obtain the Master of Science Degree in Biological Engineering Supervisors: Dr. Maria Teresa Ferreira Ces´ ario Smolders Prof. Dr. Maria Manuela Regalo da Fonseca Examination Committee Chairperson: Prof. Dr. Duarte Miguel de Franc ¸a Teixeira dos Prazeres Supervisor: Dr. Maria Teresa Ferreira Ces´ ario Smolders Member of the Committee: Dr. Maria Catarina Marques Dias de Almeida December 2016
Transcript

Xylose Upgrade

Directing the metabolism of xylose towards P(3HB), xylitol andxylonic acid in Burkholderia sacchari

Ines Maria Vicente Palolo

Thesis to obtain the Master of Science Degree in

Biological Engineering

Supervisors:

Dr. Maria Teresa Ferreira Cesario SmoldersProf. Dr. Maria Manuela Regalo da Fonseca

Examination Committee

Chairperson: Prof. Dr. Duarte Miguel de Franca Teixeira dos PrazeresSupervisor: Dr. Maria Teresa Ferreira Cesario Smolders

Member of the Committee: Dr. Maria Catarina Marques Dias de Almeida

December 2016

To my grandparents, whom I miss everyday.

Once we accept our limits we can go beyond them.Albert Einstein

Preface

The work described in this document was performed under the framework of the curricular course

Master Dissertation in order to obtain the Masters degree in Biological Engineering at Instituto Superior

Tecnico (IST).

All the research was carried out at Institute for Bioengineering and Biosciences (IBB), Department

of Bioengineering, at IST, Alameda campus, from February to November 2016, under the supervision of

the Post-Doc researcher Mª Teresa Cesario and of Professor Mª Manuela Fonseca.

This dissertation consists of a literature review within the experimental work focus, the description

of the obtained results and respective discussion. Finally, conclusions and future work prospects are

described as well.

v

Acknowledgments

First of all, I would like to thank both Professor Mª Manuela Fonseca and Doctor Mª Teresa Cesario

for giving me this amazing opportunity. The work I developed under your oversight in the past nine

months not only allowed me to obtain my Masters degree but also helped me developing my critical

sense and my engineering skills. Also, your vast knowledge in the field and your availability to advise

me and revise my work were crucial to the success of this dissertation thesis, for which I am deeply

grateful.

Now, for the rest of you, this is one of those ”once in a life time” things, and I definitely want to make

it count, so prepare yourselves because I am going to do my best to bring you all to tears!

I would like to start thanking to the engineers around: Rodrigo Raposo and Claudia Henriques, my

predecessors. Your insight knowledge and advice were precious throughout this time, although I decided

to ignore part of them and leave things until the last minute. And also Marisa Santos, who I cannot thank

enough for all the HPLC emergency responses and time spent around that cursed (but very treasured,

and also broken...) machine, and for all your support and advice during my time in the lab.

One of the major acknowledgements goes to the crew of IST’s South tower seventh floor lab: Joao

Lourenco, one of the friends college brought me, and who always cracks ”that” joke lightening up the

mood; Margarida Silva, for always telling me that everything was going to be okay; Ana Santos, for

giving life to the ”thirteen” and for your extraordinary ability to turn any serious and factual conversation

naughty (I am also glad that I can now call you an engineer as well); Ana Pfluck, for your advice and

willingness to, together with Ana S., organize all the social-cultural events in these past months; the

brazilians Darlisson D. Alexandria and Willian Birolli, who cheer up any slow work afternoon in the lab or

lunch hour with fun facts about Brazil or any other subject, and I hope I get to meet you in Sao Paulo in a

not so distant future; and also Marcel Lackner, the funniest german I met so far (and the only one), who I

should remind that offered to give me a Frankfurt city-tour. I want to thank you for all these past months,

for the lunch hours, the dinners out, the people I got to meet through you, but especially for the trips,

since I concquered more portuguese land this summer than in the last 5 years. These are memories I

will always treasure. You were my semester abroad.

I also would like to thank Carolina Gaibino, Ines Maltez and Isabel Lopes, literally my ”homegirls”, for

your support throughout these years. We witnessed each other grow into four fine young women, with

brilliant lives ahead of us. Looking back to where we started, and looking to where we are now... I am

vii

definitely proud of us.

Next in line is Joana Costa, the girl who stood by me since day three of college. We will always have

our disagreements but we just can’t get enough of each other. Thank you for your patience with my

terrible morning mood and for pushing me to get stuff done. Also, to Lia Vieira, Marlene Oliveira, Sara

Estevens, Tatiana Mangericao, and Rui Campos, you are the ”friends-for-life” college brought me, even

if for some reason we follow different paths. You’ve always stood by me and that’s enough. Aside from

my degree, you are the best I could’ve asked for during these years.

Last, but definitely not least, I would like to thank my amazing family, with all its perks, especially

my dad, Victor, my mom, Anita, and sister, Mafalda. This is your concquest too. Your unconditional

love, patience and support during all these years was everything, and without it none of this would have

never happened, and for that you have my eternal gratitude. Thank you for believing in me, sometimes

even more than I believed in myself. I cannot end ”this” without mentioning my grandparents who always

pushed me to do better and be better. Me becoming an engineer was also your dream, and I am sorry

that you’re not here to witness it... I hope I made you proud.

As strange as it sounds, I would like to finish these acknowledgements by thanking Tecnico. You

were the university I didn’t want to get into... and here we are. After seven years of engineering college,

seven years of laughter, exams, tears, group projects, barbecues and lab reports, our relatioship finally

comes to an end. These were without any doubt the best and the worst years of my life. Honestly, I

would do it all over again. It has been a hell of a ride my friend, but it’s time for me to get a life.

Cheers!

Obrigada, e um brinde a vos!

viii

Abstract

Burkholderia sacchari produces insoluble energy-storage compounds called poly-hydroxyalkanoates

(PHAs) when cultivated under unbalanced growth conditions such as excess carbon source and limita-

tion of an essential nutrient. PHAs are biodegradable and biocompatible thermoplastics and thus a

suitable alternative for petroleum-based plastics. Moreover, when cultivated on D-xylose, this strain

is able to produce D-Xylitol and D-xylonic acid. Given the capability of B. sacchari to produce differ-

ent commercially interesting compounds from D-xylose, its potential was explored within the biorefinery

concept.

Opposite to what has been published based on glucose, nitrogen limiting conditions allowed for

higher polymer production compared to phosphorus limited medium. By decreasing the oxygen supply

and thus imposing a double limitation, P(3HB) production was enhanced, achieving yields of 0.07 g

P(3HB)/g xylose and 0.12 g P(3HB)/g xylose in P- and N-limited media, respectively. Based on these

results, improvement of P(3HB) production using a xylose-rich feed (ratio xylose-glucose of 8:1) and

simulating the conditions of an assay carried out previously in our laboratory was sought. By limiting

the oxygen supply to the bioreactor, a yield of 0.15 g P(3HB)/g total sugars, and a polymer content of

30% (g P(3HB)/g CDW) were attained in conditions where no polymer production had been previously

observed.

In short, results suggest that in B. sacchari xylose-cultivations the dissolved oxygen in the medium

is a target factor to enhance P(3HB) production. Additionally, it is shown that depending on the xylose

concentration in the broth, D-xylitol and D-xylonic acid production can be promoted.

Keywords

Burkholderia sacchari, poly-3-hydroxybutyrate (P(3HB)), D-xylonic acid, D-xylitol, fed-batch cultiva-

tion, D-xylose

ix

Resumo

Quando cultivada em condicoes de crescimento desequilibradas, tais como excesso de fonte de car-

bono e limitacao de um nutriente essencial, a Burkholderia sacchari produz compostos de armazena-

mento de energia chamados poli-hidroxialcanoatos (PHAs). Estes sao termoplasticos biodegradaveis e

biocompatıveis, daı serem uma alternativa adequada aos plasticos convencionais. Quando cultivada em

D-xylose, esta estirpe tambem produz D-xilitol e acido D-xilonico. Como produz diferentes compostos

de interesse comercial, o seu potencial foi investigado num contexto de biorrefinaria.

Ao contrario do que tem sido publicado para culturas em glucose, condicoes limitantes de azoto

permitiram obter producoes de polımero mais elevadas do que em meio limitado por fosforo. Dimin-

uindo o fornecimento de oxigenio e, como tal, impondo uma dupla limitacao, a producao de P(3HB) e

melhorada, atingindo-se rendimentos de 0.07 g P(3HB)/g xilose e 0.12 g P(3HB)/g xilose em meios lim-

itados por P e por N, respectivamente. Com base nestes resultados, foi possıvel melhorar a producao

de P(3HB) usando uma alimentacao mista rica em xilose (racio de 8:1), simulando ensaios laboratoriais

efectuados anteriormente. Limitanto a disponibilidade de oxigenio no meio, a producao de polımero

e promovida, atingindo-se um rendimento de 0.15 g P(3HB)/g acucares totais, correspondendo a um

conteudo celular de 30% de polımero (g P(3HB)/g CDW).

Em suma, os resultados obtidos sugerem que em culturas de B. sacchari em xilose, o oxigenio

dissolvido no meio e um factor crucial na melhoria da producao de P(3HB), e tambem que a producao

de D-xilitol e acido D-xilonico e promovida conforme a concentracao de xilose no meio.

Palavras Chave

Burkholderia sacchari, poli-3-hidroxibutirato (P(3HB)), acido D-xilonico, D-xilitol, cultura em fed-batch,

D-xilose

xi

Contents

Preface v

Acknowledgements vii

Abstract ix

Resumo xi

List of Figures xvii

List of Tables xxi

Abbreviations xxiii

1 Introduction 1

1.1 Context . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2

1.2 Objectives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4

2 State of the Art 5

2.1 The Biorefinery Concept . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6

2.2 Xylose: the major component of the renewable feedstock lignocellulose used as sole

carbon source . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2.2.1 Lignocellulosic materials . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7

2.2.2 Xylose metabolic network . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8

2.3 Polyhydroxyalkanoates: a route towards biobased plastics . . . . . . . . . . . . . . . . . . 12

2.3.1 Polyhydroxyalkanoates (PHAs) . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12

2.3.1.A PHAs synthesis, chemical structure and physical properties . . . . . . . . 12

2.3.1.B Approaches for P(3HB) production and recovery . . . . . . . . . . . . . . 13

2.3.2 Biodegradability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15

2.3.3 Applications and commerciallisation . . . . . . . . . . . . . . . . . . . . . . . . . . 16

2.4 Xylitol and xylonic acid: value-added products from the microbial conversion of xylose-rich

lignocellulosic feedstock . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

2.4.1 Xylitol: an artificial sweetener with beneficial properties . . . . . . . . . . . . . . . 19

2.4.1.A Applications and commerciallisation . . . . . . . . . . . . . . . . . . . . . 20

xiii

2.4.1.B Approaches for xylitol production . . . . . . . . . . . . . . . . . . . . . . . 20

2.4.2 Xylonic acid . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

2.4.2.A Applications and commerciallisation . . . . . . . . . . . . . . . . . . . . . 24

2.4.2.B Approaches for xylonic acid production . . . . . . . . . . . . . . . . . . . 24

3 Aim of Studies: directing Burkholderia sacchari metabolism towards P(3HB), xylitol and

xylonic acid production 29

4 Materials and Methods 33

4.1 Microorganism . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

4.2 Strain storage and preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

4.3 Culture media . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

4.3.1 Seeding medium . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

4.3.1.A Oligoelements solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

4.3.2 Cultivation medium to impose limitation by phosphorus . . . . . . . . . . . . . . . 35

4.3.3 Cultivation medium to impose limitation by nitrogen . . . . . . . . . . . . . . . . . . 35

4.4 Carbon sources . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

4.4.1 Feed solution . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

4.5 Culture conditions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

4.5.1 Shake flask assays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

4.5.1.A Inoculum preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

4.5.1.B Shake flask cultivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

4.5.1.C Culture sampling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

4.5.2 Fed-batch assays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

4.5.2.A Inoculum preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

4.5.2.B Fed-batch cultivation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 37

4.5.2.C Culture sampling . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

4.6 Analytical methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

4.6.1 Optical density measurements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

4.6.2 Cell dry weight determination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38

4.6.3 Carbon sources, xylonic acid, xylitol and phosphate determinations . . . . . . . . 38

4.6.3.A Sample preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

4.6.3.B Calibration curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

4.6.4 P(3HB) determination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39

4.6.4.A Sample preparation: acidic methanolysis . . . . . . . . . . . . . . . . . . 40

4.6.4.B Calibration curve . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

4.6.5 Ammonium quantification . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

4.6.5.A Sample preparation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

4.6.5.B Calibration curves . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40

4.6.6 Overall yield and productivity calculations . . . . . . . . . . . . . . . . . . . . . . . 41

xiv

5 Results and Discussion 43

5.1 Shake flask assays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

5.1.1 Xylose as sole carbon source for B. sacchari cultivations . . . . . . . . . . . . . . 44

5.1.2 Influence of the C-source used for inocula growth . . . . . . . . . . . . . . . . . . . 45

5.1.3 Influence of YE presence in seeding medium . . . . . . . . . . . . . . . . . . . . . 48

5.1.4 Influence of inhibitory substrate concentrations . . . . . . . . . . . . . . . . . . . . 50

5.1.5 Influence of pH . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

5.2 Bioreactor assays . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

5.2.1 Influence of pH in B. sacchari batch cultivations . . . . . . . . . . . . . . . . . . . 52

5.2.2 Influence of the limiting nutrient in the medium used for fed-batch cultivations . . . 54

5.2.3 Influence of the dissolved oxygen availability in the culture broth . . . . . . . . . . 57

5.2.4 Cultivations towards xylitol production: influence of inhibitory xylose concentrations 61

5.2.5 Three-stage continuous bioreactors series: third bioreactor optimisation approach 63

5.2.6 B. sacchari cultivation strategies towards xylonic acid, xylitol and P(3HB) production 65

6 Conclusions and future prospects 67

Bibliography 71

Appendix A Substrates, by-products of xylose metabolism, and ammonium quantification A-1

A.1 Carbon sources, xylonic acid, xylitol and phosphate determinations . . . . . . . . . . . . . A-2

A.2 P(3HB) determination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A-4

A.3 Ammonium determination . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A-4

xv

List of Figures

2.1 Biorefinery concept schematics, integrated in the agro-biofuel-biomaterial-biopower cycle.

Based on Ragauskas (2006) and SIADEB (2010). . . . . . . . . . . . . . . . . . . . . . . 6

2.2 Sugarcanes, a raw material for sugar production. . . . . . . . . . . . . . . . . . . . . . . . 7

2.3 Sugarcane bagasse, after sugarcane processing. . . . . . . . . . . . . . . . . . . . . . . . 7

2.4 Molasses, an industrial by-product of sugar production. . . . . . . . . . . . . . . . . . . . 7

2.5 Xylose metabolism for B. sacchari, based on literature review (Jeffries, 1983; Lam et al.,

1980; Moat et al., 2002; Gottschalk, 1986; Sudesh et al., 2000; Radek et al., 2014;

Kanehisa et al., 2016; Ogata et al., 1999; Raposo et al., 2017). The enzimatic ac-

tivites are abbreviated as follows: xylose isomerase (XI); xylulokinase (XK); xylose re-

ductase (XR); xylitol dehydrogenase (XOHDH); xylose dehydrogenase (XDH); xylonolac-

tonase (XL); ribulose-phosphate 3-epimerase (RPE); transketolase (TKL); transaldolase

(TAL); phosphoketolase (PKL); ribulose-phosphate isomerase (RPI); glyceraldehyde-3-

phosphate dehydrogenase (GlyPDH); 3-phosphoglycerate kinase (PGK); enolase (EL);

pyruvate kinase (PK); pyruvate dehydrogenase complex (PDH); acetate kinase (AK);

pyrophosphate-acetate phosphotransferase (PAP); acetyl-CoA synthase (ACS); β-kethiolase

(PhaA); acetoacetyl-CoA reductase (PhaB); PHA synthase (PhaC); KDPG aldolase (KDPGA);

glucose-6-phosphate dehydrogenase (GluPDH); phosphogluconate dehydrogenase (GNDH);

phosphogluconate dehydratase (GNDY); hexokinase (HK). . . . . . . . . . . . . . . . . . 9

2.6 Two alternative metabolic routes from D-xylose to α-ketoglutarate in C. glutamicum (Radek

et al., 2014). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10

2.7 PHAs synthesis in bacteria using hydroxyacyl-CoA thioesters as precursor, where the

pendant group (R) varies from methyl (C1), to tridecyl (C13) (Verlinden et al., 2007). . . . . 13

2.8 Life cycle of PHAs: these will biodegrade to CO2, H2O, humic matter and biomass. New

agricultural crops, using nutrients from compost and fixing CO2, will produce building

blocks, monomers and polymers (Gross, 2002; Verlinden et al., 2007). . . . . . . . . . . . 15

2.9 D-Xylitol structural formula. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

2.10 Cristals of xylitol. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19

2.11 Chewing gums: the major industrial application of xylitol. . . . . . . . . . . . . . . . . . . . 19

2.12 D-Xylonic acid strutural formula. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24

xvii

3.1 TEM image of B. sacchari containing 70% (dry weight) of P(3HB) (Cesario and de Almeida,

2015). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30

4.1 Experimental set-up for the fed-batch cultivations. Inocula grown on glucose (Gluc), is

transferred to the bioreactor where it is cultivated on xylose (Xyl) with pH and temperature

control, and online DO measurement. Periodic sampling takes place for biomass, sugars

and organic acids, and P(3HB) determination. . . . . . . . . . . . . . . . . . . . . . . . . . 37

5.1 B. sacchari shake flask cultivation data for growth and production in seeding medium,

supplemented with 20 gL-1 of xylose as main carbon source, using an inoculum cultivated

on glucose. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44

5.2 Plot of equation 5.2 using experimental data from the exponential phase of the shake flask

assay represented in figure 5.1. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 45

5.3 Experimental set-up to study of the influence of the sugar used to grow the inoculum:

glucose (Gluc), xylose (Xyl) and sucrose (Suc), on the growth and production performance

of cultures grown on 20 gL-1 of xylose. Periodic sampling for pH, biomass, sugars and

organic acids, and P(3HB) determination. . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

5.4 B. sacchari shake flask cultivation data for growth and production in SM, supplemented

with 20 gL-1 of xylose as main carbon source, and an inoculum grown on glucose. . . . . 47

5.5 B. sacchari shake flask cultivation data for growth and production in SM, supplemented

with 20 gL-1 of xylose as main carbon source, and an inoculum grown on xylose. . . . . . 47

5.6 B. sacchari shake flask cultivation data for growth and production in SM, supplemented

with 20 gL-1 of xylose as main carbon source, using an inoculum grown on sucrose. . . . 47

5.7 Experimental set-up for the study of the influence of yeast extract (YE) presence in the

shake flask cultivation seeding medium. Inoculum grown on glucose (Gluc) cultivated

in complete seeding medium (SM) and in seeding medium without YE (SM-YE), sup-

plemented with 20 gL-1 of xylose (Xyl). Periodic sampling for pH, biomass, sugars and

organic acids, and P(3HB) determination. . . . . . . . . . . . . . . . . . . . . . . . . . . . 49

5.8 B. sacchari glucose grown inoculum cultivated in shake flask, in complete SM supple-

mented with 20 gL-1 of xylose as main carbon source. . . . . . . . . . . . . . . . . . . . . 49

5.9 B. sacchari glucose grown inoculum cultivated in shake flask with 20 gL-1 of xylose as

main carbon source, in seeding medium without YE. . . . . . . . . . . . . . . . . . . . . . 49

5.10 Experimental set-up for the study of xylitol production in shake flask cultivations. Inocu-

lum grown on glucose cultivated in seeding medium (SM) supplemented with 50 gL-1 of

xylose (Xyl). Periodic sampling for pH, biomass, sugars and organic acids, and P(3HB)

determination. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50

5.11 Glucose grown inoculum cultivated in seeding medium, supplemented with high concen-

trations of substrate (50 gL-1 of xylose). . . . . . . . . . . . . . . . . . . . . . . . . . . . . 51

5.12 B. sacchari pH profile during shake flask cultivation in seeding medium, supplemented

with 20 gL-1 of xylose. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52

xviii

5.13 Data obtained in B. sacchari batch cultivation A, using seeding medium supplemented

with 50 gL-1 of xylose, with pH controlled at 6.8 and 20% of DO. Equations A.1 to A.6

were used as calibrations curves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

5.14 Data obtained in B. sacchari batch cultivation B, using seeding medium supplemented

with 50 gL-1 of xylose, with 20% of DO and without pH control. Equations A.1 to A.6 were

used as calibrations curves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53

5.15 Data obtained in B. sacchari fed-batch cultivation C, using P limited medium supple-

mented with 30 gL-1 of xylose as main carbon source, at pH 6.8 and 20% of DO. A 600

gL-1 solution of xylose was used as feed. Equations A.7 to A.12 were used as calibration

curves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

5.16 Data obtained in B. sacchari fed-batch cultivation D, using N limited medium supple-

mented with 30 gL-1 of xylose as main carbon source, at pH 6.8 and 20% of DO. A 600

gL-1 solution of xylose was used as feed. Equations A.1 to A.6, and A.14 were used as

calibrations curves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55

5.17 Data obtained in B. sacchari fed-batch cultivation E, using P limiting medium supple-

mented with 30 gL-1 of xylose as main carbon source, at pH 6.8 and 1% of DO. A 600

gL-1 solution of xylose was used as feed. Equations A.1 to A.6 were used as calibrations

curves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

5.18 Data obtained in B. sacchari fed-batch cultivation F, using P limiting medium supple-

mented with 30 gL-1 of xylose as main carbon source, at pH 6.8 and 1% of DO. A 600

gL-1 solution of xylose was used as feed. Equations A.7 to A.12 were used as calibrations

curves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 58

5.19 Data obtained in B. sacchari fed-batch cultivation G, using N limiting medium supple-

mented with 30 gL-1 of xylose as main carbon source, at pH 6.8 and 1% of DO. Manual

pulses of a 600 gL-1 solution of xylose were added during cultivation as feed. Equations

A.7 to A.12, and A.15 were used as calibrations curves. . . . . . . . . . . . . . . . . . . . 59

5.20 Data obtained in B. sacchari fed-batch cultivation H, using N limiting medium supple-

mented with 30 gL-1 of xylose as main carbon source, at pH 6.8 and a set point for DO of

1%. A total volume of 316 mL of a 600 gL-1 xylose solution was used as feed. Equations

A.7 to A.12, and A.16 were used as calibrations curves. . . . . . . . . . . . . . . . . . . . 59

5.21 Data obtained in B. sacchari fed-batch cultivation I, using P-limited cultivation medium

supplemented with 30 gL-1 of xylose as main carbon source, at pH 6.8 and 20% of DO. A

total volume of 198 mL of a 600 gL-1 xylose solution was used as feed. Equations A.7 to

A.12 were used as calibrations curves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

5.22 Data obtained in B. sacchari fed-batch cultivation J, using P-limited cultivation medium

supplemented with 30 gL-1 of xylose as main carbon source, at pH 6.8 and 20% of DO. A

total volume of 236 mL of a 600 gL-1 xylose solution was used as feed. Equations A.7 to

A.12 were used as calibrations curves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 62

xix

5.23 Data obtained in B. sacchari fed-batch cultivation K, using P limiting medium supple-

mented with 30 gL-1 of xylose, at pH 6.8 and 20% of DO. A mixed solution of 600 gL-1

of xylose and 75 gL-1 of glucose was used as feed. Equations A.7 to A.12 were used as

calibrations curves. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64

5.24 Data obtained in B. sacchari fed-batch cultivation L, using P-limited medium supple-

mented with 30 gL-1 of xylose, at pH 6.8 and 1% of DO. A total volume of 26 mL of a

mixed solution of 600 gL-1 of xylose and 75 gL-1 of glucose was used as feed. Equations

A.7 to A.12 were used as calibrations curves. . . . . . . . . . . . . . . . . . . . . . . . . . 64

A.1 HPLC chromatogram for RI detector. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . A-3

A.2 HPLC chromatogram for UV-visible detector. . . . . . . . . . . . . . . . . . . . . . . . . . A-3

xx

List of Tables

2.1 P(3HB) chemical and physical properties. Adapted from Verlinden et al. (2007); Lee

(1996); Henriques (2015). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

2.2 Screening of strains reported to produce P(3HB) from D-xylose catabolism (Cesario and

de Almeida, 2015). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 18

2.3 Screening of microorganisms for D-xylitol production by biological conversion of D-xylose. 23

2.4 Screening of microorganisms for D-xylonic acid (XylAc) production by biological conver-

sion of D-xylose. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27

4.1 Seeding medium composition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 34

4.2 Oligo elements solution composition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

4.3 Phosphorus limited medium composition. . . . . . . . . . . . . . . . . . . . . . . . . . . . 35

4.4 Nitrogen limited medium composition. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

5.1 Overall yields of xylonic acid (XylAc) and P(3HB) in shake flask cultivations using 20 gL-1

of xylose as sole carbon source, and inocula grown in three different sugars: glucose,

xylose and sucrose. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

5.2 Overall yields of xylonic acid (XylAc) and P(3HB) in shake flask cultivations using seeding

medium with and without yeast extract (YE). . . . . . . . . . . . . . . . . . . . . . . . . . . 49

5.3 Overall yields of xylonic acid (XylAc), xylitol (Xylt) and P(3HB) in shake flask cultivations

with inhibitory concentrations of xylose (50 gL-1). . . . . . . . . . . . . . . . . . . . . . . . 50

5.4 Overall yields and productivities of xylonic acid (XylAc), xylitol (Xylt) and P(3HB) in biore-

actor batch cultivations, using seeding medium supplemented with 50 gL-1 of xylose, with

(A) and without (B) pH control, and 20% saturation of DO in the medium. . . . . . . . . . 53

5.5 Overall yields and productivities of xylonic acid (XylAc), xylitol (Xylt) and P(3HB) in fed-

batch cultivations C and D, carried out in an essential nutrient limiting medium supple-

mented with 30 gL-1 of xylose, at pH 6.8 and 20% of DO in the medium. . . . . . . . . . . 56

5.6 Overall yields and productivities of xylonic acid (XylAc), xylitol (Xylt) and P(3HB) in fed-

batch cultivations using phosphorus (cultivations E and F) and nitrogen (cultivations G

and H) limited medium supplemented with 30 gL-1 of xylose, at pH 6.8 and 1% saturation

of DO. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60

xxi

5.7 Overall yields and productivities of xylonic acid (XylAc), xylitol (Xylt) and P(3HB) in the

bioreactor fed-batch cultivations using a phosphorus limited medium, supplemented with

30 gL-1 of xylose, at pH 6.8 and 20% saturation of DO. . . . . . . . . . . . . . . . . . . . . 62

5.8 Overall yields and productivities of xylonic acid (XylAc) and P(3HB) in the third bioreactor

simulating fed-batch cultivation using medium limited by phosphorous, supplemented with

30 gL-1 of xylose, at pH 6.8 with different oxygen saturations. Note that yield calculations

were performed in terms of total of sugars consumed. . . . . . . . . . . . . . . . . . . . . 64

5.9 Summary of the results obtained for B. sacchari bioreactor cultivations strategies towards

xylonic acid (XylAc), xylitol (Xylt) and/or P(3HB) production throughout this work. Residual

biomass (Xr) and maximum yield obtained for each metabolite under the various operation

conditions studied are shown. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66

xxii

Abbreviations

A. niger Aspergillus niger

acetoacetyl-CoA acetoacetyl-coenzyme A

acetyl-CoA acetyl-coenzyme A

ATP Adenosine Triphosphate

B. sacchari Burkholderia sacchari

C. crescentus Caulobacter crescentus

C. necator Cupriavidus necator

CDW cell dry weight

CCR carbon catabolite repression

DO dissolved oxygen

DOE Department of Energy

E. coli Escherichia coli

ED Entner-Doudoroff

EMP Embden-Meyerhoff-Parnas

FAD Flavin Adenine Dinucleotide

GC Gas Chromatography

Gluc glucose

GNDY phosphogluconate dehydratase

GO glucose oxidase

H. boliviensis Halomonas boliviensis

HPLC High Performance Liquid Chromatography

IS internal standard

xxiii

K. lactis Kluyveromyces lactis

K. pneumoniae Klebsiella pneumoniae

KC Krebs cycle

KDPG 2-keto-3-deoxy-6-phosphogluconate

KDPGA KDPG aldolase

kLa volumetric oxygen transfer coefficient

LC lignocellulosic biomass

LCF lignocellulosic feedstock

µmax maximum specific cell growth rate

mcl-PHA medium chain length PHA

NAD Nicotinamide Adenine Dinucleotide (oxidized form)

NADH Nicotinamide Adenine Dinucleotide (reduced form)

NAD(P) Nicotinamide Adenine Dinucleotide Phosphate (oxidized form)

NAD(P)H Nicotinamide Adenine Dinucleotide Phosphate (reduced form)

OD optical density

OTR oxygen transfer rate

P product

P. kudriavzevii Prichia kudriavzevii

PDH pyruvate dehydrogenase complex

PEP phosphoenolpyruvate

PHA poly-hydroxyalkanoate

PhaA β-ketothiolase

PhaB acetoacetyl-CoA reductase

PhaC poly-hydroxyalkanoate synthase

PHB poly-hydroxybutyrate

P(3HB) poly-3-hydroxybutyrate

PHV poly-hydroxyvalerate

PK pyruvate kinase

xxiv

PPP Pentose Phosphate

RI refraction index

RPE ribulose-phosphate 3-epimerase

RPI ribulose-phosphate isomerase

S substrate

S. cerevisiae Saccharomyces cerevisiae

scl-PHA small chain length PHA

SM seeding medium

STR stirred-tank reactor

Suc sucrose

t time

T. reesei Trichoderma reesei

TAL transaldolase

TCA tricarboxylic acid cycle

TEM Transmission Electronic Microscope

TKL transketolase

UV Ultraviolet

X biomass

XDH xylose dehydrogenase

XDY xylonate dehydratase

XOHDH xylitol dehydrogenase

Xyl D-xylose

XylAc D-xylonic acid

Xylt D-xylitol

XI xylose isomerase

XK xylulokinase

Xr residual biomass

XR xylose reductase

xxv

YE yeast extract

ZW zero waste

xxvi

1Introduction

1

1.1 Context

Currently, the world population faces two major problems as regards the energy demand and waste

generation. Not only both industrialised and developing countries require an increasingly amount of

energy difficult to fulfil sustainably, but also the energy demand is projected to grow more than 50%

by 2025 (Ragauskas, 2006), and, morevover, the amount of residues generated is growing. Population

growth, booming economy, rapid urbanization, and the rise in living standards have accelerated the solid

waste generation in the world (Minghua et al., 2009; Song et al., 2015). In fact, human population is

expected to increase up to 9.7 billions by 2050 (United Nations Department of Economic and Social

Affairs Population Division, 2015). In 2011, the global solid waste volume was estimated about 11 billion

t per year, and per capita as approximately 1.74 t per year. Along with the large solid waste generation,

an enormous amount of natural resources are depleted everyday (Song et al., 2015).

In the past years, our society has been driven towards sustainability as it becomes more conscious

about the environment. The zero waste (ZW) concept has been highly embraced once it stimulates sus-

tainable production and consumption, optimum recycling and resource recovery, while restricting mass

incineration and landfilling. The Zero Waste International Alliance defines ZW as ”a goal that is both

pragmatic and visionary, to guide people to emulate sustainable natural cycles, where all discarded ma-

terials are resources for others to use”. Also, adopting this strategy means ”designing and managing

products and processes to reduce the volume and toxicity of waste and materials, conserve and recover

all resources, and not burn or bury them” (Zero Waste International Alliance, 2009). However, according

to Zaman (2015), in today’s society, it is not possible to achieve a 100% diversion rate1 in production,

consumption and waste management systems, i.e. it is still not possible to conduct all the waste gen-

erated towards reuse, recycling or composting. Actually, it is necessary a universal transformation of

existing extraction, production, marketing, consumption, management and treatment systems (Zaman,

2015).

Within the ZW concept, the material flow is circular, which means no materials are wasted or un-

derused. If at their end of lives products cannot be reused or repaired, they can instead be recycled

or recovered from the waste stream and used as inputs, substituting the demand for the extraction of

natural resources (Song et al., 2015). Shifting society’s dependence from petroleum-based to renew-

able biomass-based resources is generally viewed as key to the development of a sustainable industrial

society, energy independence, and to the effective management of greenhouse gas emissions (PNNL

and NREL, 2004; Ragauskas, 2006; FitzPatrick et al., 2010).

Biomass represents an abundant carbon-neutral renewable resource for the production of bioenergy

and biomaterials through biorefinery manufacturing technologies. This biorefinery vision will contribute

to sustainability not only by its inherent dependence on sustainable bioresources, but also by recycling

waste. An integrated biorefinery is an approach that optimises the use of biomass for the production of

biofuels, bioenergy, and biomaterials for both short- and long-term sustainability (Ragauskas, 2006).

In 2004, the US Department of Energy (DOE) identified 30 chemicals, derived from the conversion1Diversion rate is defined as the percentage of total waste that is diverted from disposal at permitted landfills and transformation

facilities such as incineration, and instead is directed to reduction, reuse, recycling and composting programs (CalRecycle, 2012).

2

of biomass, which could be used as building block chemicals2 in a bio-based economy (PNNL and

NREL, 2004). Biomass carbohydrates will provide a viable route to products such as alcohols, carboxylic

acids, and esters. Bio-based feedstocks are already having an impact on some practical applications,

including solvents, plastics, lubricants, and fragrances. However, these commercially viable processes

require purified feedstocks. The major impediment to biomass use is the development of methods

to separate, deconstruct, refine, and transform it into chemicals and fuels (Ragauskas, 2006). Since

chemical production requires low volumes of biomass compared to fuel production, there has been an

increasing research interest in the value of bio-sourced materials and effective use of biomass feedstock

(FitzPatrick et al., 2010). The lignocellulosic biomass (LC), the worldwide most abundant renewable raw

material, comprises different fractions such as carbohydrates, proteins, and fats that can be converted

to value-added products, fuels, and chemicals through the implementation of the biorefinery concept.

LC is a promising choice as raw material due to the high worldwide availability, in consequence of the

waste generated by agricultural and forestry processes, to its renewable nature, and to being a carbon

source which is a nonedible feedstock (Cesario and de Almeida, 2015).

Burkholderia sacchari is a Gram-negative poly-hydroxyalkanoate-accumulating bacterium isolated

from the soil of a sugar-cane plantation in Brazil (Bramer et al., 2001). These insoluble energy-storage

compounds called poly-hydroxyalkanoates (PHAs) are produced by many microbial strains under unbal-

anced growth conditions such as presence of excess carbon source or limitation of an essential nutrient

(Lee, 1996; Verlinden et al., 2007). Due to its properties, PHAs are biodegradable and biocompatible

thermoplastics, thus the interest from industry in this strain, capable of metabolising different carbon

sources (namely glucose, xylose, organic acids, etc.), and reach high cell densities, for conventional

petroleum-derived products replacement (Bramer et al., 2001; Cesario et al., 2014; Alexandrino et al.,

2015). Several studies found that B. sacchari accumulates up to 70% of the cell dry weight as poly-3-

hydroxybutyrate (P(3HB)) with different carbon sources (Bramer et al., 2001; Silva et al., 2004; Cesario

et al., 2014; Raposo et al., 2017).

Previous work developed by Henriques (2015) with B. sacchari towards P(3HB) production in a three-

stage continuous system demonstrated low P(3HB) productivities since xylose, one of the main compo-

nents of lignocellulosic hydrolysates, was not consumed for polymer production. Raposo et al. (2017)

carried other studies with this bacterial strain which revealed production of xylitol and xylonic acid, two

value-added chemical compounds, as by-products of bacterial growth on xylose. These studies mo-

tivated further research regarding polymer, xylonic acid and xylitol production using solely xylose as

substrate, and the growth conditions that promote each metabolite formation.

Cultivations of B. sacchari using xylose as main carbon source were performed to study the growth

conditions that promote P(3HB), xylonic acid, and/or xylitol production in both shake flasks and fed-

batch bioreactor cultivations, and the corresponding metabolic pathways. In this context, cultivation

requirements such as the amount of carbon source, nutrient limitations, or oxygen supply were assessed

throughout this work. Furthermore, efforts were made in order to enhance polymer production in the third

bioreactor of the continuous series, by unveiling one possible target factor.2Building block chemicals are molecules that can be converted into several secondary chemicals and intermediates, and, in

turn, into a broad range of different downstream applications.

3

1.2 Objectives

The main goal of this experimental work was the study of the growth conditions that promote poly-

3-hydroxybutyrate, xylonic acid and/or xylitol productions in fed-batch cultivations of B. sacchari using

xylose as main carbon source. The three major objectives for this dissertation are the study of:

1. the conditions that promote xylose metabolism towards P(3HB) production.

2. the bacteria’s potential to produce other important metabolites such as xylitol and xylonic acid.

3. B. sacchari xylose metabolic network.

4

2State of the Art

5

2.1 The Biorefinery Concept

The current energy demand and waste generation crisis, along with the environmental impact caused

by natural non-renewable resources depletion in developing industries and nations, has led towards

a bio-based economy, shifting society’s dependence away from petroleum to renewable biomass re-

sources (Ragauskas, 2006). While the economy of energy can be based on various alternative raw

materials (such as wind, sun, water, biomass, nuclear fission and fusion), the material economy of sub-

stances mainly depends on biomass, in particular plant biomass (Kamm and Kamm, 2004). An economy

based on innovative and cost-efficient use of biomass for the production of both bio-based products and

bioenergy should be driven by well-developed integrated biorefinning systems (de Jong and Jungmeier,

2015).

The IEA Bioenergy Task 42 (van Ree and van Zeeland, 2014) defined biorefinery as ”the sustainable

processing of biomass into a spectrum of marketable products and energy”. Thus, a biorefinery can re-

late to a facility, a process, a plant, or even a cluster of facilities. Biorefineries are expected to contribute

to an increased competitiveness and wealth of the countries by responding to the need for supplying a

wide range of bio-based products and energy in an economically, socially, and environmentally sustain-

able manner (de Jong and Jungmeier, 2015).

Figure 2.1: Biorefinery concept schematics, integrated in the agro-biofuel-biomaterial-biopower cycle. Based on Ragauskas(2006) and SIADEB (2010).

The majority of biological raw materials are produced in agriculture, forestry, and by microbial sys-

tems. Waste biomass and biomass of nature and landscape cultivation are valuable organic reservoirs

of raw material and must be used in accordance to their organic composition. Because of low cost, plen-

tiful supply, and amenability to biotechnology, carbohydrates appear likely to be the dominant source

of feedstocks for biocommodity processing. Advantages over cellulosic materials include much larger

ultimate supply, lower purchase cost and lower anticipated transfer cost, less erosity, and lower inputs

of chemicals and energy required for production. The lignocellulosic feedstock (LCF)-biorefinery will

probably be the most successful. On the one hand the raw material situation is optimal, on the other

hand conversion products have a good position in the traditional petrochemical as well as in the future

6

Figure 2.2: Sugarcanes, a raw materialfor sugar production.

Figure 2.3: Sugarcane bagasse, aftersugarcane processing.

Figure 2.4: Molasses, an industrial by-product of sugar production.

bio-based product market (Kamm and Kamm, 2007).

2.2 Xylose: the major component of the renewable feedstock lig-nocellulose used as sole carbon source

The most frequently reported factor that influences the price of PHA is the cost of the carbon source,

whose selection should not focus only on the market prices, but also on the availability and global price

consistency. Inexpensive carbon sources such as agricultural wastes and industrial by-products may

incur additional costs due to pre-treatment steps, extended cultivation times, and purification. Although

simple carbon sources, such as sugar and starch from crops, seem to be superior to complex carbon

sources, they are also a primary source of human food and animal feed (Chanprateep, 2010).

2.2.1 Lignocellulosic materials

Lignocellulosic materials consist of three primary chemical fractions or precursors: (i) hemicellu-

lose/polyases, a sugar polymer of predominantly pentoses; (ii) cellulose, a glucose-polymer; and (iii)

lignin, a polymer of phenols (Kamm and Kamm, 2007). Cellulose and hemicellulose are polysaccha-

rides embedded in a complex lignin matrix, which gives the plant structural support, impermeability, and

resistance against microbial attack and oxidative stress (Cesario and de Almeida, 2015).

Biological conversion of lignocellulosic materials into value-added building block chemicals and fuels

usually involves several sequential steps: lignocellulose pretreatment/fractionation, enzymatic cellulose

hydrolysis to fermentable sugars (e.g., glucose and xylose), and fermentation (Cesario and de Almeida,

2015). In order to be used by microorganisms, lignocellulosic waste must undergo hydrolysis, from

which different sugars are released from the cellulosic and hemicellulosic fractions and, depending on

the biomass nature, different hexose/pentose ratios are attained (Raposo et al., 2017). Since agricultural

wastes like wheat straw, rice straw and corn stover are rich in xylose and, generally, have a lower lignin

content, they make a perfect fit as feedstock. For instance, molasses, represented in figure 2.4, is

one potentially inexpensive carbon source for PHA production. However, its value been increasing to

unprecedented levels at both origin and destination (Chanprateep, 2010).

Although xylose is a major component of lignocellulose and the second most abundant sugar in

nature, efficient use is still a technical barrier due to the strong physical and chemical construction of

lignocellulose (Venkateswar Rao et al., 2016). Therefore, the effective biological conversion of lignocel-

7

lulosic hydrolysates is restricted by the ability of the strain used to consume pentoses (Raposo et al.,

2017).

The major difficulty associated with the use of sugar mixtures and thus with lignocellulosic hy-

drolysates is the carbon catabolite repression (CCR), a regulatory mechanism that prevents the expres-

sion of the genes needed for other carbon sources catabolism, usually pentoses, while the substrate

that enables the fastest growth (usually glucose) is present. In Escherichia coli cultivations carried with

sugar mixtures as carbon sources, diauxic growth is observed, i.e. a sequential consumption of the

sugar takes place resulting in a series of exponential growth phases separated by lag phases. The CCR

regulation mechanism varies depending on the bacterial species (Raposo et al., 2017). For instance, in

Burkholderia sacchari cultivations Cesario et al. (2014) observed that in media containing glucose and

citrate, the uptake of glucose started upon citrate exhaustion. Moreover, CCR has been reported in this

strain with glucose/xylose mixtures containing 10 gL-1 of glucose plus 10 gL-1 of xylose (Cesario et al.,

2014).

2.2.2 Xylose metabolic network

For the purpose of understanding the fundamentals of this work, an overview of the metabolic re-

actions involved in xylose metabolism by bacteria was carried out. The network described in figure

2.5 represents the proposed xylose metabolism for Burkholderia sacchari (B. sacchari), and was con-

structed based on literature review (Jeffries, 1983; Lam et al., 1980; Moat et al., 2002; Gottschalk, 1986;

Sudesh et al., 2000; Radek et al., 2014; Kanehisa et al., 2016; Ogata et al., 1999; Raposo et al., 2017).

Glucose conversion reactions are represented as well, however these will not be detailed in the follow-

ing text. It is important to note that the network is far from comprising all the possible metabolic pathways.

Bacteria generally employ active transport mechanism for the uptake of sugars and other nutrients.

Active transport mechanisms are mediated by carrier proteins, and hence exhibit the properties of sat-

urability, substrate specificity, and specific inhibition. This process occurs at the expense of metabolic

energy, but sugars can be transported against a concentration gradient. The metabolic energy required

can be provided by establishing a membrane potential as in the chemiosmotic energization mechanism,

by the hydrolysis of Adenosine Triphosphate (ATP) as in the direct energization mechanism, or by the

transfer of phosphate from phosphoenolpyruvate (PEP) to the sugar substrate as in the group translo-

cation mechanism (Jeffries, 1983).

In Escherichia coli , the transport of D-xylose across the cell membrane is linked to the movement of

protons, as evidenced by a rise in pH in the extracellular medium upon addition of D-xylose. In other

words, D-xylose transport is energized by a chemiosmotic mechanism in which protons are transported

together across the cell membrane, and not by directly energized or PEP-phosphotransferase mech-

anisms (Jeffries, 1983). Chemiosmotic-driven transport systems accomplish movement of a molecule

across the membrane at the expense of a previously established ion gradient such as a proton motive

force or a sodium motive force. There are three basic types: symport, which involves the simultaneous

transport of two substrates in the same direction by a single carrier; antiport, that also involves simul-

8

Figure 2.5: Xylose metabolism for B. sacchari, based on literature review (Jeffries, 1983; Lam et al., 1980; Moat et al., 2002;Gottschalk, 1986; Sudesh et al., 2000; Radek et al., 2014; Kanehisa et al., 2016; Ogata et al., 1999; Raposo et al., 2017).The enzimatic activites are abbreviated as follows: xylose isomerase (XI); xylulokinase (XK); xylose reductase (XR); xylitoldehydrogenase (XOHDH); xylose dehydrogenase (XDH); xylonolactonase (XL); ribulose-phosphate 3-epimerase (RPE); trans-ketolase (TKL); transaldolase (TAL); phosphoketolase (PKL); ribulose-phosphate isomerase (RPI); glyceraldehyde-3-phosphatedehydrogenase (GlyPDH); 3-phosphoglycerate kinase (PGK); enolase (EL); pyruvate kinase (PK); pyruvate dehydrogenase com-plex (PDH); acetate kinase (AK); pyrophosphate-acetate phosphotransferase (PAP); acetyl-CoA synthase (ACS); β-kethiolase(PhaA); acetoacetyl-CoA reductase (PhaB); PHA synthase (PhaC); KDPG aldolase (KDPGA); glucose-6-phosphate dehydroge-nase (GluPDH); phosphogluconate dehydrogenase (GNDH); phosphogluconate dehydratase (GNDY); hexokinase (HK).

taneous transport of two like-charged compounds by a common carrier but in opposite directions; and

uniport, which occurs when movement of a substrate is independent of any coupled ion (Moat et al.,

2002). Lam et al. (1980) observation indicate that xylose transport is energized by a proton-motive

force. When xylose was added to anaerobic energy-depleted cells, an alkaline pH change expected for

a sugar-H+ symport system occurred. Contrary to phosphotransferase systems, which give an acid shift,

or directly energized systems that yield no pH change (Lam et al., 1980).

There are at least four different pathways for the catabolism of D-xylose described in literature (Ogata

et al., 1999; Kanehisa et al., 2016), namely an Oxido-reductase pathway, an Isomerase pathway, and

9

two oxidative pathways, called Weimberg and Dahms. Prokaryotes typically use an isomerase pathway,

and the two oxidative pathways (Cherix et al., 2014). The Weimberg pathway (Weimberg, 1961) repre-

sented in figure 2.6 is an oxidative pathway where D-xylose is oxidized to D-xylonolactone by a xylose

dehydrogenase (XDH) followed by a lactonase to hydrolyse the lactone to D-xylonic acid. A xylonate

dehydratase (XDY) is splitting off a water molecule resulting in 2-keto-3-desoxyxylonate. A second de-

hydratase forms the 2-keto-glutarate semialdehyde which is subsequently oxidized to 2-ketoglutarate.

For instance, Frost (2008) reported the ability of E. coli to catabolise D-xylonic acid, involving a NAD-

dependent D-XDY. The Dahms pathway (Stephen Dahms, 1974) starts as the Weimberg pathway but

the 2-keto-3-desoxy-xylonate is split by an aldolase to pyruvate and glycoaldehyde.

Raposo et al. (2017) reported the accumulation of D-xylonic acid in the broth when xylose is present

as substrate in B. sacchari cultivations, suggesting the presence of the Weimberg pathway in this strain’s

metabolism.

Figure 2.6: Two alternative metabolic routes from D-xyloseto α-ketoglutarate in C. glutamicum (Radek et al., 2014).

In order to identify which of the pathways men-

tioned are enconded in B. sacchari genome, Cherix

et al. (2014) conducted an in silico study using

the enzymes present in all four xylose pathways.

The isomerase pathway was found to be the xy-

lose metabolic pathway used by this bacterium, since

xylose isomerase (XI), xylulokinase (XK) and ABC

xylose transporter were genes present in its genome.

Some of the genes involved in the Weimberg path-

way were also found to have homologous genes in

B. sacchari . However, it was noted the absence

of the gene for 2-keto-3-desoxy-D-xylonate dehy-

dratase, which may be the reason for the reported

accumulation of D-xylonic acid, since it cannot be

converted into α-ketoglutarate semialdehyde. The

products from each of the pathways are directed to

the central metabolism of the bacteria through a dif-

ferent way, towards the Pentose Phosphate (PPP), in the case of the isomerase, and to the tricarboxylic

acid cycle (TCA) cyle in Weimberg’s. The genes of the Oxidoreductase and Dahm’s pathways were not

found in the B. sacchari genome (Cherix et al., 2014).

Once inside the bacterial cell, D-xylose is first converted to D-xylulose, a reversible isomerisation

reaction catalysed by D-XI, and then phosphorylated to D-xylulose-5-phosphate by D-XK. Both enzymes

are specifically induced by D-xylose in bacteria. On the other hand, the reduction of D-xylose to xylitol is

catalysed by XR, whereas the reoxidation of xylitol to D-xylulose is catalysed by xylitol dehydrogenase

(XOHDH). This reaction is readily reversible and NADH oxidation occurs, since this enzyme showed no

activity with NAD(P)H (Jeffries, 1983). D-xylose is required for the induction of xylose reductase (XR)

and XOHDH activities, hence xylitol formation does not occur in the absence of this pentose. Together

10

with aeration, D-xylose concentration affects xylitol formation the most (Winkelhausen and Kuzmanova,

1998).

Bacteria have been shown to possess both XR for the production of xylitol and NAD-dependent

pentitol dehydrogenases for the assimilation of pentitols, whereas the production of these sugar alcohols

is essentially a function of the NAD(P)H-specific aldolase reductase (Jeffries, 1983). Nevertheless,

Bramer et al. (2001) reported that B. sacchari does not assimilate xylitol as carbon source.

D-xylose degradation continues towards PPP pathway, a series of reversible transaldolase (TAL)

and transketolase (TKL) reactions, which primary functions are to provide NAD(P)H, which is used

primarily for reducing power in biosynthetic reactions, and two very important biosynthetic precursors,

sedoheptulose-7-phosphate and erythrose-4-phosphate, which are essential for aromatic amino acids

(Moat et al., 2002). The pathway consists of an oxidative phase that converts hexose phosphates to pen-

tose phosphates, and a non-oxidative phase that converts pentose phosphates back to hexose phos-

phates (Jeffries, 1983). Rybulose-5-phosphate is maintained in equilibrium with rybose-5-phosphate

and xylulose-5-phosphate by the action of ribulose-phosphate isomerase (RPI) and ribulose-phosphate

3-epimerase (RPE), respectively (Moat et al., 2002).

Carbon exits from the sugar phosphate pool by at least three and sometimes four routes: D-fructose-

6-phosphate and glyceraldehyde-3-phosphate can enter the Embden-Meyerhoff-Parnas (EMP) pathway;

D-rybose-5-phosphate is used for nucleotide synthesis; D-erythrose-4-phosphate is the starting point for

the shikimic acid pathway leading to the synthesis of aromatic amino acids; and, in some organisms,

D-xylulose-5-phosphate can form glyceraldehyde-3-phosphate plus acetyl phosphate through the action

of pyruvate kinase (PK) (Jeffries, 1983).

After pyruvate formation, the next step towards energy production for the cell is the oxidative de-

carboxylation of pyruvate to acetyl-coenzyme A (acetyl-CoA) by the pyruvate dehydrogenase com-

plex (PDH), and then the oxidation of acetyl-CoA via the TCA, commonly known as Krebs cycle (KC).

The cycle carries out the oxidation of the acetyl moiety of acetyl-CoA to CO2 with transfer of the re-

ducing equivalents to Nicotinamide Adenine Dinucleotide (oxidized form) (NAD)+, NAD(P)+, and FAD

(Gottschalk, 1986).

As regards the P(3HB) formation, two acetyl-CoA moities are combined to form acetoacetyl-CoA by

a β-ketothiolase (PhaA). The product then undergoes reduction by an NAD(P)H-dependent reductase,

acetoacetyl-CoA reductase (PhaB), which produces the (R)-isomer of 3-hydroxybutyryl-CoA. Finally, the

poly-hydroxyalkanoate synthase (PhaC) polimerises the 3-hydroxybutyryl-CoA to poly-hydroxybutyrate

(PHB) inclusions through the formation of ester bonds between the hydroxyl groups of the monomers,

with the liberation of coenzyme-A. PhaC shows a broad substrate specificity and thefore a wide variety of

monomers can be polimerised. The most important factors that determine the type of PHA constituents

is the carbon source, and only (R)-isomers are accepted as substrates for the polymerising enzyme due

to its stereospecificity (Sudesh et al., 2000; Verlinden et al., 2007).

According to (Gomez et al., 2014), B. sacchari showed flux exclusively in Entner-Doudoroff (ED)

pathway, replacing the first reactions in the EMP pathway. The two key enzymes for ED pathway are

11

6-phosphogluconate dehydratase (GNDY), which catalyses dehydration of 6-phosphogluconate to form

2-keto-3-deoxy-6-phosphogluconate (KDPG), and KDPGA, which cleaves KDPG to pyruvate and glyc-

eraldehyde 3-phosphate, the latter being further catabolised through the EMP pathway and TCA cycle.

2.3 Polyhydroxyalkanoates: a route towards biobased plastics

Plastic materials have become an integral part of contemporary life on account of many desirable

properties including durability and resistance to degradation. However, throughout the past decades,

these nondegradable plastics have accumulated in the environment (Lee, 1996), leading to catastrophic

consequences, namely groundwater and land pollution, and animal life endangerment.

Recently, the society have become more conscious about this subject, and efforts have been made

in several fronts towards an effective solid waste management and sustainability. These concerns have

created a renewed interest in biologically derived polymers, ignored for a long time due to high pro-

duction cost compared to petrochemical derived plastics. During recent years, intensive research has

investigated the bacterial production of PHAs and a great effort is underway to improve this procedure.

Nonetheless, the PHA production price is still far above the price of conventional plastics (Verlinden

et al., 2007).

2.3.1 Polyhydroxyalkanoates (PHAs)

In order to find alternative materials to oil-based plastics, researchers have developed fully biodegrad-

able plastics, such as PHAs which show similar material properties to polypropylene. Other advantages

of these materials over petrochemical plastics are that they are natural, renewable and biocompatible

(Verlinden et al., 2007). Another important feature of PHAs is that they are degraded naturally by mi-

croorganisms which enable carbon dioxide and organic compound recycling in the ecosystem, providing

a buffer to climate change (Chanprateep, 2010).

Bacteria synthesise and accumulate PHAs as carbon and energy storage materials or as a sink for

redundant reducing power under the condition of limiting nutrients, such as nitrogen, phosphorus or

oxygen, in the presence of generous carbon supplies (Lee, 1996; Sudesh et al., 2000). Also, by poly-

merising soluble intermediates into these insoluble PHAs granules, the cell does not undergo alterations

of its osmotic state (Verlinden et al., 2007), and, once accumulated, the polymer serves as both carbon

and energy source during starvation (Sudesh et al., 2000).

2.3.1.A PHAs synthesis, chemical structure and physical properties

The first poly-hydroxyalkanoate to be discovered thus the most widely studied, was described by

Lemoigne in 1926 as an unknown material in the form of a homopolyester of the 3-hydroxybutyric acids

called poly-hydroxybutyrate (PHB) (Suriyamongkol et al., 2007). As described in section 2.2.2, within the

bacterial metabolism carbon substrates are converted into hydroxyl-CoA thioesters, which are converted

into PHB by three biosynthetic enzymes, as seen in figure 2.7 (Verlinden et al., 2007).

12

Figure 2.7: PHAs synthesis in bacteria using hydroxyacyl-CoA thioesters as precursor, where the pendant group (R) varies frommethyl (C1), to tridecyl (C13) (Verlinden et al., 2007).

Table 2.1: P(3HB) chemical and physical properties. Adapted from Verlinden et al. (2007); Lee (1996); Henriques (2015).

Parameter Value

Decomposition temperature (°C) 246.3

Melting temperature (°C) 177

Glass transition temperature(°C) 2

Crystallinity (%) 60

Tensile strengh (MPa) 40

Extension to break (%) 5

Other properties

Non-toxic, good UV light resistance, relative resistance to hydrolytic degradation, water

insoluble, good oxygen impermeability, sinks in water facilitating anaerobic degradation

The array of aplications for PHAs stems from the enormous variation possible in the length and

composition of the side chains. For instance, PHB and poly-hydroxyvalerate (PHV) form a class of

PHAs typically referred to as small chain length PHAs (scl-PHAs), whereas PHAs composed of C6 to

C16 3-hydroxy fatty acids are known as medium chain length PHA (mcl-PHA). Copolymers of PHB are

formed when mixed substrates are used, and these will be random copolymers unless substrates are

alternated overtime. PHB homopolymer is a highly crystalline thermoplastic, stiff, but brittle substance,

and when spun into fibres it behaves as a hard-elastic material (Verlinden et al., 2007). These and other

characteristics that make PHB a suitable alternative for petroleum-based plastics are summarised in

table 2.1.

2.3.1.B Approaches for P(3HB) production and recovery

PHAs are synthesised by many living organisms including plants, yeasts and bacteria. Plants can

only cope with low yields of PHA production (<10% (w/w)), due to the negative effects higher levels of

polymer have on its growth and development. In contrast, bacteria can accumulate PHAs up to 90%

(w/w) of of the cell dry weight. Bacterial fermentation of Cupriavidus necator is by far the most exten-

sively studied strain for PHAs production, and it seems to be the most cost-effective process. A few

important other strains capable of producing PHAs by way of fermentation processes include: Bacillus

spp., Alcaligenes spp., Pseudomonas spp., Aeromonas hydrophila, Rhodopseudomonas palustris, Es-

cherichia coli , Burkholderia sacchari and Halomonas boliviensis (Verlinden et al., 2007). According to

Suriyamongkol et al. (2007), currently, more than 300 microorganisms are reported to synthesise PHAs.

Since the scope of this work is xylose biological conversion, in table 2.2 is given an overview of the

strains reported to be able to metabolise this pentose and produce P(3HB).

13

One of the main current research focus regarding polymer production is productivity maximisation,

i.e. the goal is to obtain the highest density of products of interest in a reaction volume within the shortest

time (Ienczak et al., 2013). Several approaches and operation modes from batch to continuous, with or

without cell recycling, have been studied and reviewed throughout the last decades (Verlinden et al.,

2007; Ienczak et al., 2013). Despite the broad spectrum of strains able to accumulate PHAs, research

groups have struggled to optimise polymer production when inexpensive carbon sources are used as

substrate (Chanprateep, 2010).

In fed-batch cultivations cells are grown under a batch regime usually until close to the end of the ex-

ponential phase, and then a solution of substrates is fed to the bioreactor, without removal of culture broth

(Ienczak et al., 2013). The fermentative production of PHAs is normally operated as a two-stage fed-

batch process: an initial growth phase in nutritionally enriched medium, which yields sufficient biomass,

followed by a product formation phase in a nitrogen-depleted medium. Single fed-batch fermentation

that are nitrogen limited lead to low amounts of polymer, since there is not enough accumulation of

biomass (Verlinden et al., 2007).

Industrial PHA production only becomes economically viable when high cell densities and high poly-

mer production are achieved at low cost. PHAs’ productivity is proportional to the residual biomass pro-

duction once this polymer accumulates inside the cells. Therefore, processes with high-cell-densities,

i.e. higher residual biomass concentration, are preferred because they allow for culture volume re-

duction, for lower production costs and lower investment in equipment compared to low-cell-density

processes (Ienczak et al., 2013).

Chanprateep (2010) cited academic research groups that have focused on other approaches for

PHB production optimisation. For instance, mathematical models have been developed to control the

fermentation conditions in order to maximise PHB production, strategies such as maintaining a constant

nutrient concentration to promote polymer accumulation, and on-line feeding strategies for fed-batch

cultivations have also been studied (Chanprateep, 2010).

It has been reported that resorting to mixed bacterial cultures for PHAs production can improve the

efficiency of the fermentation, fomenting the decrease of the PHAs’ cost due to lower sterility, equip-

ment and control requirements, and the ability to utilise a wide range of cheap substrates (including

industrial and agricultural wastes), and therefore increase their market potential (Verlinden et al., 2007;

Chanprateep, 2010).

PHAs’ production in mixed cultures is induced by an intracellular limitation, through a process referred

to as feast and famine. In this strategy, cells are exposed to consecutive periods of substrate accessibility

(”feast”) and then unvailability (”famine”), altering the phisiology of bacteria. These conditions allow for

the selection of an enriched culture with a high stable capacity of PHA production. A sudden increase of

carbon substrate supply causes the cells to change their physiology again, however, as PHA-synthesis

requires less adaptation than growth, the culture channels the substrate towards polymer accumulation

instead of cell growth (Dias et al., 2006; Verlinden et al., 2007).

PHA production in bacteria and yeast requires growth under sterile conditions, incurring the process

with high costs. In contrast, in plant systems the production of polymer is considerably less expensive

14

since it only relies on water, soil nutrients, atmospheric CO2 and sunlight. Regarding the life cycle of

PHAs described in figure 2.8, it is also a more environmentaly friendly process as plants use phospho-

synthetically fixed CO2 and water to generate the bioplastic, which after disposal is degraded back to

CO2 and water. In fact, eucaryotic systems have been studied as PHAs production hosts as an alterna-

tive to bacteria to gather information about how these pathways work and how they can be incorporated

into plants (Suriyamongkol et al., 2007).

Genetic engineering has been proved very useful in bacterial metabolism optimisation towards poly-

mer production. In fact, according to Ienczak et al. (2013), generally recombinant microorganisms have

showed more competitiveness for industrial processes than wild-type strains due to the absence of de-

polymerase enzyme, the absence of the necessity of nutrient limitation, and the capacity to produce

higher P(3HB) content whithin a short time period. However, recombinant microorganisms have shown

instability for industrial processes as well, requiring antibiotics and inductors for expressing the genes of

interest (Ienczak et al., 2013).

To recover the polymer formed during cultivation, bacterial cells containing the PHA are in the first

place separated from the medium by centrifugation. Then, organic solvents, such as acetone, chloro-

form, methylene chloride or dichloroethane, have been used to extract the intracellular product. How-

ever, in industrial processes it is necessary a large amount of solvent what turns the procedure both

economically and environmentally unappealing. Nevertheless, when it comes to medical applications,

the solvent extraction is a good method since the resulting PHA presents high purity. As an alternative,

aqueous enzymatic procedures, treatments with ammonia, or digestion with sodium hypochlorite and

surfactants have been proposed. More recently, supercritical fluid disruption, dissolved-air flotation, and

selective dissolution of cell biomass for the recovery of PHAs have also been studied (Verlinden et al.,

2007).

2.3.2 Biodegradability

Figure 2.8: Life cycle of PHAs: these will biodegrade toCO2, H2O, humic matter and biomass. New agriculturalcrops, using nutrients from compost and fixing CO2, willproduce building blocks, monomers and polymers (Gross,2002; Verlinden et al., 2007).

Conventional polymers such as polyethylene and

polypropylene persist in the environment for many

years after disposal. Additionally to being inappro-

priate for applications in which plastics are used for

short time periods and then disposed, plastics are of-

ten soiled by food and other biological substances,

making physical recycling impractical and generally

undesirable (Gross, 2002).

In his review, Lee (1996) cited a number of aero-

bic and anaerobic PHA-degrading bacteria and fungi

isolated from various environments, namely from soil,

activated sludge, sea and also lake water, by using

PHA hydrolases and PHA depolymerases. The en-

zymes’ activity and hence the rate of PHA biodegra-

15

dation, may vary according to the composition of the polymer and the environmental conditions, such

as the microbial population and the temperature. Moreover, the presence of UV light can accelerate the

process. The end products of PHA degradation in aerobic environments are carbon dioxide and water,

while methane is also produced in anaerobic conditions (Verlinden et al., 2007; Lee, 1996).

As shown in figure 2.8, the synthesis and biodegradation of PHAs are totally compatible with the

carbon-cycle once its fermentative production uses agricultural feeds such as sugars and fatty acids as

carbon and energy sources. Despite the fact that PHAs are renewable compounds, which is one of the

most notable features of these polymers, studies into their life cycle showed that more energy would be

needed, from crop growing to moulding the final product, than in the life cycle of conventional plastics.

Nonetheless, the fermentation process to make PHA is far from optimised (Verlinden et al., 2007), and

research regarding polymer production using lignocellulosic wastes as raw materials is underway.

2.3.3 Applications and commerciallisation

The majority of expected applications of PHAs, whether used in pure form or as additives to oil-

derived plastics such as polyethylene, are as replacements for petrochemical polymers. Target markets

for the biopolymers include packaging materials (trash bags, wrappings, loose-fill foam, food containers,

film wrapping, laminated paper), disposable nonwovens (engineered fabrics) and hygiene products (di-

aper back sheets, cotton swabs), consumer goods (fast-food tableware, containers, egg cartons, razor

handles, toys), and agricultural tools (encapsulation of seeds and of fertilizers for slow release, mulch

films, planters). PHAs have also been processed into toners for printing applications and adhesives for

coating applications (Gross, 2002; Verlinden et al., 2007).

Since PHAs are a product of cellular metabolism, and also 3-hydroxy butyric acid (the product of

degradation) is present in blood, these polymers have an ideal biocompatibility, feature that makes them

suitable for medical applications. In pure form or mixed with other materials, PHAs are used in sutures,

repair patches, orthopedic pins, adhesion barriers, stents, nerve guides and bone marrow scaffolds.

Polymer implants for targeted drug delivery can be made out of PHAs as well. However, due to the high

level of specification for plastics used in the human body, the medical PHA has to be free of bacterial

endotoxins and consequently there are high requirements for the extraction and purification methods

(Verlinden et al., 2007).

PHA are undoubtedly one of the potential candidates for replacing petroleum-based plastics. How-

ever, the cost of PHB is much higher than that of other bio-based polymers due to high raw material costs,

small production volumes, and high processing costs, particularly purification. Only few biodegradable

plastics made of PHA are available in the market (Mirel™, Biocycle™, Biomer™, etc.) (Chanprateep,

2010). In 2011, the prices for PHAs were in the range of 3.7-4.5 C per kg, while conventional polyolefins

such as polyethyleneterephalate and polysterene were in the range of 1.38-1.63 C per kg (Cesario et al.,

2014), which is still too expensive to justify the shift to the ”green” alternative.

In order to make the process economically attractive, many goals have to be addressed towards

optimisation. For instance, better recovery/purification steps and use of inexpensive substrates can

also substantially reduce the production cost, turning the fermentation process more efficient. Another

16

approach is the development of recombinant microbial strains to achieve both high substrate conversion

rate and close packing of PHAs granules in the host cell. Additionally, research to enhance the physical

properties of PHAs is also required (Verlinden et al., 2007).

17

Table 2.2: Screening of strains reported to produce P(3HB) from D-xylose catabolism (Cesario and de Almeida, 2015).

Strain Riskgroup

CDW(gL-1)

%P(3HB)(gg-1)

YP(3HB)/Xyl(gg-1)

Prodvol(gL-1h-1) Operation mode References

Burkholderia cepaciaATCC 17759 2 7.5 49 0.11

iBatch: xylose 30 gL-1 (N-limited medium)

Young et al.(1994)

Burkholderia cepacia IPT048 2 4.4 53 0.29 0.09 Batch: sugarcane bagasse hy-

drolysate 16 gL-1Silva et al.(2004)

Burkholderia sacchariIPT 101 1 5.5 58 0.26 0.07 Batch: xylose 15 gL-1 (N-

limited medium)Lopes et al.(2009)

Burkholderia sacchariLMF828 (mutant PTS-

glu+1 5.3 50 0.17 0.07 Batch: xylose 15 gL-1 Lopes et al.

(2011)

Bacillus cereus CFR06 2 1.1 35i i

Batch: xylose 2% (w/w) Halami (2008)

Bacillus sp. MA3.3 (Bacil-lus megaterium) 1 5.5 64 0.24 0.06 Batch: xylose 15 gL-1 (N-

limited medium)Lopes et al.(2009)

Escherichia coliTG1(pSYL107) 2 4.8 36

i

0.028 Batch: xylose 20 gL-1 Lee (1998)

Pseudomonas pseud-oflava ATCC 33668 1 4.0 22 0.017 1 Batch: xylose 44 molmol-1 (N-

limited medium)Bertrand et al.(1990)

Isolated bacterium strainQN271

i

4.3 29i

0.04i

Doan andNguyen (2012)

Burkholderia sacchariDSM 17165 1

6.3 44 0.24 0.08 Batch: xylose 20 gL-1 Cesario et al.(2014)

11.3 35 0.15 0.028Batch: xylose 30 gL-1; Fed-batch: xylose 600 gL-1 gL-1 (N-limited medium, 1% O2 sat.)

Present work

i Unknown

18

Figure 2.9: D-Xylitol structural formula.

Figure 2.10: Cristals of xylitol.Figure 2.11: Chewing gums: the majorindustrial application of xylitol.

2.4 Xylitol and xylonic acid: value-added products from the mi-crobial conversion of xylose-rich lignocellulosic feedstock

At the beginning of the 20th century, many industrial materials such as dyes, solvent and synthetic

fibers were made from trees and agricultural crops, being later displaced by petroleum derivatives. Due

to the energy crisis of the 1970s, a renewed interest in the synthesis of fuels and materials from biore-

sources arose. The paradigm shift from petroleum hydrocarbons to highly oxygen-functionalized, bio-

based feedstocks will create remarkable opportunities for the chemical processing industry. In fact,

bio-based feedstocks are already having an impact on some practical applications, including solvents,

plastics, lubricants, and fragrances, and biomass carbohydrates are expected to provide a viable route

to products such as alcohols, carboxylic acids, and esters (Ragauskas, 2006). The reported block chem-

icals by PNNL and NREL (2004) can be produced out of sugar via biological and chemical conversions,

and can subsequently be converted to a number of high-value biobased chemicals and/or materials

(Kamm and Kamm, 2007), such as xylitol and xylonic acid.

2.4.1 Xylitol: an artificial sweetener with beneficial properties

D-Xylitol, a five carbon sugar alcohol, occurs widely in nature, in many fruits and vegetables, but is

also a normal intermediate in human metabolism (Winkelhausen and Kuzmanova, 1998). At an industrial

scale, this artificial sweetener can be produced from the second most abundant polysaccharide, xylan

rich hemicellulose which upon hydrolysis produces xylose. D-Xylitol has attracted worldwide interest due

to its unique properties and potential. It has almost the same sweetness as sucrose, but lower energy

value (2.4 calg-1 vs. 4.0 calg-1), thus it has been used as a sugar substitute in dietary foods. It has

also found applications in pharmaceutical industries due to its properties such as anticarcinogenicity,

tooth rehardening, preventive against otitis, ear and upper respiratory infections, etc. (Chen, 2010).

Tough most of xylitol is produced chemically based on catalytic reduction of pure xylose under high

pressure and temperature using an expensive catalyst (usually Ni-catalyst), biotechnological production

of xylitol from xylose offers a better alternative in terms of energy consumption and overall process cost

(Venkateswar Rao et al., 2016).

Some of the lignocellulosic raw materials used for the production of xylitol mainly include corncobs,

wheat straw, corn stover, wheat bran, miscanthus, etc. A prerequisite step in the production of xylitol is

the pretreatment of the substrate which is essential for removal of lignin and to reduce the crystallinity

19

of the substrate. Hydrolysis of lignocellulosic biomass using acids is most widely used, during which

optimisation of various process parameters is required with a focus on xylose recovery. Detoxification

process reduces the concentration of inhibitors. Later the sugars, especially xylose will, be converted to

xylitol during fermentation and the xylitol thus produced is further concentrated, purified and crystallised

(Venkateswar Rao et al., 2016).

2.4.1.A Applications and commerciallisation

Xylitol finds a broad application either as the sole sweetener or in conjunction with other sweeteners

in the preparation of full- and reduced-energy sugarless confectionery products suitable for infants and

diabetics. Other potential applications of xylitol are food products such as bakery products, spices,

relishes, jams, jellies, marmalades and desserts, and also pharmaceuticals and oral hygiene products

(Winkelhausen and Kuzmanova, 1998).

Nowadays, consumers are inclining towards sugar free and low calorie food products due to health

and weight consciousness, increasing the demand for xylitol. The global consumption of xylitol was

approximately 160 thousand metric tons in 2013 which is equivalent to market value of US$ 670 million,

and is expected to reach 242 thousand metric tons by 2020, equivalent to US$ 1 billion. The chewing

gum industry has the largest market for xylitol consumption compared to other industries and was found

to be 80% in 2010 and estimated to consume about 163 thousand metric tons by 2020 with 60% of

global consumption (Venkateswar Rao et al., 2016). Xylitol has a relatively high value (4.5–5.$/kg for

bulk purchase by pharma/chewing gum companies, and 20$/kg in supermarkets), which makes it an

attractive proposition for commerciallisation. Therefore, over the past few decades much effort has been

devoted to the development of cost-effective and environmentally-friendly biotechnological processes by

evaluating cheaper lignocellulosic substrates (Ravella et al., 2012). Xylitol production cost varies and

depends mainly on cost of the raw material used and its transport cost which depends on feedstock

mass and location of the manufacturing plant. (Venkateswar Rao et al., 2016).

2.4.1.B Approaches for xylitol production

Although xylitol occurs in many fruits and vegetables, the content of xylitol is usually low and thus it

is very uneconomical to extract it from such sources (Chen, 2010). On a large scale, xylitol is currently

produced by chemical reduction of xylose derived mainly from wood hydrolysates. The conventional

process include (i) acid hydrolysis of the plant material, (ii) purification of the hydrolysate to either a pure

xylose solution or a pure crystalline xylose, (iii) hydrogenation of xylose to xylitol, and (iv) crystallinization

of the xylitol (Winkelhausen and Kuzmanova, 1998).

The large scale production of xylitol is attained by nickel catalysed chemical process that is based

on xylose hydrogenation, which requires purified xylose as raw substrate. This chemical reaction re-

quires high temperature and pressure, which makes this process cost intensive and energy consuming

(Venkateswar Rao et al., 2016). After the first steps that originate the xylose-rich hydrolysate, it is then

subjected to hydrogenation at 80-140°C, and hydrogen pressures up to 50 atm in the presence of metal

catalysts. The xylitol solution formed requires further purification by chromatography, and then concen-

20

tration and crystallisation of the product to obtain pure xylitol. The yield is only about 50-60% of the

xylan fraction, which incurs the process with extensive separation and purification stages, thus xylitol

production becomes expensive (Rafiqul and Mimi Sakinah, 2012). For this reason, there is a necessity

to develop an integrated economically feasible process for biotechnological conversion of lignocellulosic

materials into xylitol.

The microbial xylitol production has been greatly investigated as an alterantive to the chemical pro-

cess as it does not need extensive purification steps for xylose and is operated at low cost (Venkateswar

Rao et al., 2016), and also the downstream processing is expected to be cheaper (Winkelhausen and

Kuzmanova, 1998). Microorganisms like bacteria, yeasts ans fungi possess the ability of fermenting

commercial xylose or xylose present in hydrolysates derived from lignocellulosic residues to xylitol. Xyl-

itol production has been extensively carried out using free or immobilised cells in batch, continuous or

fed-batch process, using synthetic media or lignocellulosic substrates (Venkateswar Rao et al., 2016).

Microbial production of xylitol is influenced by various fermentation parameters, such as the car-

bon and nitrogen sources, the aeration, and the effect of pH and temperature. Aiming at an efficient

xylitol production process, optimisation of these factors, which depend on the microorganism used, is

mandatory.

According to Venkateswar Rao et al. (2016), organic nitrogen sources are more beneficial than in-

organic nitrogen sources and show a greater impact on cell growth and xylitol production. In fact, the

addition of urea to a Candida athensensis SB18 cultivation in synthetic medium enhanced the con-

version of xylose (Zhang et al., 2012). Oxygen availability in the medium, which is often expressed

in terms of volumetric oxygen transfer coefficient (kLa) and oxygen transfer rate (OTR), is an impor-

tant process parameter that plays a major role in regulating cell growth and xylitol formation, since it

affects xylose reductase (XR) and xylose dehydrogenase (XDH) enzymatic activities. Himabindu and

Gummadi (2015) studied the effect of oxygen in terms of kLa, and concluded that xylitol production is

favoured by micro-aerobic conditions, i.e. lower kLa values. Hernandez-Perez et al. (2016) studied the

effect of various sugars as co-substrate for xylitol production by Candida guilliermondii FTI20037, by

supplementing sugar cane bagasse hydrolysate containing 57 gL-1 xylose with sugars such as maltose,

sucrose and glycerol. It was found that the addition of sucrose alone favoured xylitol production, leading

to the possibility of using sugarcane molasses as nutritional supplementation of fermentation media to

supply sucrose as co-substrate and replace nutrients already exhausted (Hernandez-Perez et al., 2016).

Continuous fermentation maintains high productivities for longer periods of time by eliminating the

idle time for cleaning and sterilisation, and it shows greater steadiness in product synthesis. However,

immobilised cells are generally used in this operation mode, as they possess high cell density, improved

stability with higher productivity and can be re-used when compared to free cells (Venkateswar Rao

et al., 2016).

In recent decades, various metabolic engineering strategies have been explored to modify the en-

zymes XR and XDH in microorganisms in order to increase the production of xylitol. Strain improvement

methods, such as mutation, adaptation, and recombination, have also been employed to enhance the

21

performance of microorganisms used in fermentation processes (Venkateswar Rao et al., 2016).

Rafiqul and Mimi Sakinah (2012) reviewed an alternative promising approach for conversion of xylose

into xylitol resorting to enzymatic technology, since fermentation processes on an industrial-scale are

not feasible due to decreased productivity. During the enzymatic process, xylose was stoichiometrically

converted to xylitol with almost an equivalent consumption of NAD(P)H. After that, an almost quantitative

conversion of xylose to xylitol was achieved using NAD+ to a xylose ratio of over 1:30. At the same

time, the coenzyme was successfully regenerated and retained using a membrane reactor. About 90%

conversion of xylose to xylitol could be achieved at 35°C and pH 7.5 after a 24h reaction period (Rafiqul

and Mimi Sakinah, 2012).

A brief literature review of the microorganisms producers of xylitol is summarised in 2.3, as well as

the main cultivation conditions used in this regard.

22

Table 2.3: Screening of microorganisms for D-xylitol production by biological conversion of D-xylose.

Strain Riskgroup

YXylt/Xyl(gg-1)

tferm(h)

Prodvol(gL-1h-1)

Operation mode References

Corynebacterium sp. 1 or 2 0.69 336 0.21 Batch: xylose 150 gL-1 Yoshitake et al.(1971)

Enterobacter sp. 2 0.49 96 0.34 Batch: xylose 100 gL-1 Yoshitake et al.(1973)

Candida athensensisSB18 2

0.83 181 1.15 Batch: xylose 250 gL-1

Zhang et al.(2012)0.87 264 0.97 Fed-batch: xylose 300 gL-1

0.81 102 0.98 Fed-batch: xylose 123.42±0.05 gL-1

Candida tropicallis ASMIII

2 0.93 120 1.08 Batch: xylose 200 gL-1 (40% O2 sat.) Lopez et al.(2004)

Candida tropicallisKCTC 10457 2

0.83 47 3.53 Batch: xylose 200 gL-1 (1 vvm) Kwon et al.(2006)0.9 48 4.88 Fed-batch: xylose 260 gL-1 (1 vvm)

Candida guilliermondiiFTI 20037 2

0.71 72 0.56 Batch: xylose 54.5 gL-1 from rice strawhydrolysate

Roberto et al.(1995)

0.7 55 0.64 Batch: xylose 61 gL-1 from sugar canebagasse hydrolysate

Debaryomyces nepalen-sis NCYC 3413

i 0.83 108 0.83 Batch: sucrose 200 gL-1; Fed-batch:xylose 50 gL-1 (0.5 vvm)

Himabinduand Gummadi(2015)

Burkholderia sacchariDSM 17165

1 0.056 130 0.018 Batch: xylose 50 gL-1 (seedingmedium, 20% O2 sat.)

Present work

i Unknown

23

2.4.2 Xylonic acid

Figure 2.12: D-Xylonic acid struturalformula.

Aldonic acids are sugar acids produced by oxidation at the C-1 of

the corresponding aldoses, namely xylose, glucose, mannose, arabi-

nose, and galactose (Pezzotti and Therisod, 2006). The aldonic acids

belong to a class of polyhydroxyl acids that are currently generating

considerable interest and have significant economic potential as chela-

tors, buffers and preservatives, and they may also serve as platform

reagents for the synthesis of other useful chemicals. Together with

other polyhydroxy carboxylic acids, aldonic acids are known to be produced by electrochemical or chem-

ical oxidation. However, due to their pollution effects and processing cost this method is inaccessible for

the commercial production (Zhou et al., 2015).

The production of aldonic acid chemicals by using the lignocellulose as a cost-competitive raw ma-

terial is promising for biomass refinery. However, many barriers will appear due to the presence of

inhibitors, unknown colloidal matter, and various aldoses in the hydrolysates used as the starting mate-

rial (Zhou et al., 2015).

D-xylose can be converted into D-xylonic acid by chemical or biological routes, such as microbial cul-

tivations or resorting to purified enzymes. In the next pages, D-xylonic acid applications and production

approaches will be assessed, as well as a screening of the microorganisms reported to produce this

by-product of the xylose catabolism.

2.4.2.A Applications and commerciallisation

Polyhydroxy acids and mostly sugar acids have been proved to be useful in industrial, medical and

agricultural applications, but especially as sequestering agents for metal ions (Pezzotti and Therisod,

2006).

D-xylonic acid, an oxidation product of xylose, has been identified as among the top 30 value-added

chemicals from biomass (PNNL and NREL, 2004), and is one of the most promising potential products

from xylose, since the five carbon structure of xylose remains intact. The main applications of D-xylonic

acid are conversion to lactones and esters, and it has potential as a sequestrant due to its similarity

to D-gluconic acid in molecular structure (Buchert et al., 1988; Wang et al., 2016). The conversion of

xylose to xylonic acid is the first step of the 1,2,4-butanetriol biosynthesis, which is used to synthesise

1,2,4-butanetriol trinitrate, an energetic material (explosive) superior to nitroglycerin (Niu et al., 2003). In

addition, D-xylonic acid can also be used as concrete additive which improves concrete dispersion (Liu

et al., 2012).

2.4.2.B Approaches for xylonic acid production

D-Xylose can be converted into D-xylonic acid by chemical or biological routes. The most frequently

used methods for aldonic acids preparation use stoichiometric amounts of bromine, copper or silver

hydroxydes in totally non-ecological processes (Pezzotti and Therisod, 2006), namely: oxidation of the

parent aldose; synthesis from lower aldoses; degrative oxidation of aldoses; epimerization of other

24

aldonic acids; and various other methods (De Lederkremer and Marino, 2004). However, these will not

be described since this work focuses mainly on biological conversion of xylose. For further consultation,

De Lederkremer and Marino (2004) performed an extensive review on this matter. On the other hand,

microbial conversion of D-xylose to D-xylonate with bacteria, including Acetobacter sp., Pseudomonas

aeruginosa, Pseudomonas fragi, and Gluconobacter oxydans, has been well described (Wang et al.,

2016). Besides native microorganisms, the xylonic acid synthesis pathway has been constructed in

many recombinant microorganisms, by overexpressing a NAD+-dependent D-xylose dehydrogenase.

Wang et al. (2016) investigated D-xylonic acid production using D-xylose and hydrolysed bamboo as

substrates. Xylose is a suitable substrate of the glucose dehydrogenase in the glucose oxidation path-

way, an important pathway for glucose catabolism in Klebsiella pneumoniae. When used as substrate,

xylose led to xylonic acid accumulation in the broth, a process dependent upon acidic conditions. Using

the hydrolysate of bamboo as substrate, mixture of 33 gL-1 gluconic acid and 14 gL-1 xylonic acid was

produced by the recombinant K. pneumoniae. In fed-batch fermentation, 103 gL-1 of xylonic acid was

produced after a cultivation period of 79 h, with a conversion ratio of 1.11 gg-1 (Wang et al., 2016).

A D-xylonic acid-producing Kluyveromyces lactis strain was constructed by expressing a NAD(P)+-

dependent D-xylose dehydrogenase (XDH) from Trichoderma reesei , and 19 gL-1 of D-xylonic acid was

produced from 40 gL-1 of D-xylose (Nygard et al., 2011). Liu et al. (2012) reported the construction of

an engineered E. coli to produce D-xylonic acid. To do so, the native pathway for D-xylose catabolism in

E. coli W3110 was blocked by disrupting xylose isomerase and xylulokinase genes. The native pathway

for xylonic acid catabolism was also blocked by disrupting two genes both encoding xylonate dehy-

dratase. Through the introduction of a NAD+-dependent XDH from Caulobacter crescentus, a D-xylonic

acid producing E. coli was constructed (Liu et al., 2012). This XDH gene was also expressed in Sac-

charomyces cerevisiae (Toivari et al., 2012), and Prichia kudriavzevii (Toivari et al., 2013).

Resorting to a different approach, Pezzotti and Therisod (2006) reported a very simple procedure for

several aldonic acids synthesis from the corresponding aldoses oxidation, catalysed by glucose oxidase

(GO), an enzyme extremely specific for D-glucose. GO from Aspergillus niger is a robust enzyme

produced on an industrial scale and marketed in large quantities at reasonable prices from several

companies. Oxidation was performed by simple dissolution of the sugar in water, followed by addition

of glucose oxidase under vigorous agitation, and catalase to decompose the hydrogen peroxide formed

as a byproduct, which is known to have a deleterious effect on GO. The aldonic acid formed during

oxidation was continuously neutralised with NaOH, maintaning medium’s pH during the reaction. The

volume of NaOH added is directly proportional to the amount of sugar being oxidised. After filtration,

a pure aldonic acid was obtained, free from the enzymes and from any unreacted substrate. Using

A. niger glucose oxidase as the catalyst, 1.98 gL-1 xylonic acid was produced from xylose with a yield of

90%. Temperature and pure oxygen atmosphere conditions are very important variables that influence

reaction rates, and immobilisation of GO and catalase should be considered for large scale synthesis

(Pezzotti and Therisod, 2006).

Despite all the ”green” production approaches towards xylonic acid production, so far, no commercial

production of D-xylonic acid has been established. Reasons are either because bacteria strains produce

25

many other oxidising enzymes resulting in the conversion of other sugars present in lignocellulosic hy-

drolysates, or due to the low D-xylonic acid accumulation rate and yield that engineered yeast strains

have. In addition, the high cost of peptone and/or yeast extract media as nitrogen sources is generally

uneconomical for industrial scale productions (Liu et al., 2012).

Table 2.4 summarises the screening of microorganisms reported to produce D-xylonic acid through

cultivation processes, as well as the main conditions used to do so.

26

Table 2.4: Screening of microorganisms for D-xylonic acid (XylAc) production by biological conversion of D-xylose.

Strain Riskgroup

YXylAc/Xyl(gg-1)

tferm(h)

Prodvol(gL-1h-1) Operation mode References

Pseudomonas fragiATCC 4973 1 1.08 115 1.4 Batch: xylose 150 gL-1 Buchert and Vi-

ikari (1988)

Gluconobacteroxydans 1

1.10 42 2.5 Batch: xylose 100 gL-1 + glucose 5gL-1

Buchert et al.(1988)

0.90 24 1.29 Batch: xylose 30 gL-1 Zhou et al.(2015)

Kluyveromyces lactis 1 0.60 1 0.158 Batch: xylose 40 gL-1 Fed-batch: xy-lose 40 gL-1

Nygard et al.(2011)

Saccharomyces cere-visiae 1 0.80

i

0.23 Fed-batch: xylose 23 gL-1 (glucoseand ethanol as co-substrates)

Toivari et al.(2012)

Pichia kudriavzevii 2 1.0ii 170 1.4 Fed-batch: xylose 171 gL-1 (24.7 gL-1

glucose as co-substrate)Toivari et al.(2013)

Klebsiella pneumoniae 2 1.11 79 1.30 Fed-batch: xylose 500 gL-1 Wang et al.(2016)

Escherichia coli W3110 1 0.98 36 1.09 Fed-batch: xylose 40 gL-1 + glucose 10gL-1 (xylose 60 gL-1 as feed) Liu et al. (2012)

Burkholderia sacchariDSM 17165 1

1.2 41 2.2 Batch: xylose 30 gL-1; Fed-batch: xy-lose 600 gL-1 (20% O2 sat.)

Raposo et al.(2017)

1.07iii 141 0.73Batch: xylose 30 gL-1; Fed-batch: xy-lose 600 gL-1 (N-limited medium, 20%O2 sat.)

Present work

i Unknownii g per g of substrate consumediii simultaneous P(3HB) production

27

28

3Aim of Studies: directing Burkholderiasacchari metabolism towards P(3HB),

xylitol and xylonic acid production

29

As mentioned previously, biobased plastics such as PHAs are currently far more expensive than

petrochemically based plastics and are therefore used mostly in applications that conventional plastics

cannot perform, such as medical applications (Verlinden et al., 2007). One of the major drawbacks that

limit the success for industrial production via fermentation and commercialization of PHAs is the cost

of the substrate, fact that motivated extensive research in PHA production using raw materials, such

as agricultural wastes. The choice of a renewable and inexpensive carbon source is the key for a PHA

production process economically feasible (Cesario and de Almeida, 2015).

Lignocellulosic hydrolysates are rich in C5 sugars. Cesario and de Almeida (2015) described a

limited number of strains able to metabolise these substrates and convert them into PHAs, in which

Burkholderia sacchari , the bacterial strain selected to carry this experimental work, is included. These

are preferred both to increase the total carbon uptake by the cells and to avoid pentose accumulation in

the broth in fed-batch assays, therefore inhibitory concentrations.

B. sacchari was isolated from the soil of a sugar cane plantation in Brazil. The cells of strain IPT 101

are Gram-negative, rod-shaped and motile due to the presence of several polar flagella. When plated

on nutrient broth medium (Difco), white and opaque colonies are formed as result of PHA accumulation.

The optimum growth temperature range is 28-30°C, growing well between 25 and 37 °C. B. sacchari

IPT 101 demonstrated to oxidise several carbohydrates, namely glucose, xylose, and sucrose, with-

out assimilating xylitol (Bramer et al., 2001). A Transmission Electronic Microscope (TEM) image of

B. sacchari DSM 17165 cells with PHA granules is represented in figure 3.1.

Figure 3.1: TEM image of B. sacchari containing 70% (dry weight) of P(3HB) (Cesario and de Almeida, 2015).

Silva et al. (2004) studied P(3HB) production and accumulation on xylose, on xylose+glucose, and on

bagasse hydrolysates. B. sacchari IPT 101 showed promising results on xylose, accumulating 58% of

P(3HB). However, CCR prevented an efficient metabolisation of pentoses in sugar mixtures composed

of glucose, xylose, and arabinose by this wild-type strain. Through CCR the cells select from a mixture

of carbon sources the one that allows for the highest gowth rate, by inhibiting cell’s transport capacity of

other carbon sources, enzymatic activity, and related gene expression (Cesario and de Almeida, 2015).

In bacteria, the biochemical mechanisms of D-xylose metabolism are quite different from those of

D-glucose. Whereas the latter is metabolised by the EMP pathway, D-xylose metabolism proceeds by

way of the PPP pathway. Following transport into the cell, D-xylose is either isomerized or reduced, then

reoxidised to form D-xylulose. This pentulose is then phosphorylated, isomerized, and rearranged to

30

form a metabolic pool of phosphorylated 3-, 4-, 5-, 6-, and 7-carbon sugars. These intermediates can

exit the PPP to be used by other metabolic pathways through the formation of nucleic acids, aromatic

amino acids, lipids, and other metabolic end products (Jeffries, 1983).

Henriques (2015) described a three stage continuous bioreactor cascade mode for P(3HB) produc-

tion using simulated lignocellulosic hydrolysates as carbon source. Within this continuous system, the

goal was to achieve high cell densities in the first bioreactor, through a batch phase where phosphorus

was initially provided in excess, to accumulate P(3HB) in the second fed-batch bioreactor limited by phos-

phorus, by consuming preferably the glucose in the feed, and to convert the sugars present in the xylose-

rich medium in the third bioreactor, to increase the polymer yield. However, results showed low polymer

productivities and conversion yields attained, namely 0.90 gP(3HB)L-1h-1 and 0.10 gP(3HB)/gGluc+Xyl, re-

spectively, for a dilution rate of 0.1 h-1.

Since studies in a two-stage continuous system demonstraded higher polymer productivities and

yields on glucose, 2.21 gP(3HB)L-1h-1 and 0.31 gP(3HB)/gGluc, respectively, also for a dilution rate of 0.1 h-1,

the problem was thought to be the fact that B. sacchari was not able to convert the xylose present in

the medium towards polymer production because of the metabolic pathways triggered by glucose in the

second bioreactor. On the other hand, xylitol production took place in the second and third bioreactors,

with a maximum overall volumetric productivity of 0.28 gXyltL-1h-1 and a maximum overall xylitol yield of

0.13 gXylt/gXyl. It is thus, assumed that xylose accumulation and, therefore, high xylose concentrations

in the broth leads to xylitol production.

Aiming to find the reasons for this setback on continuous P(3HB) production using lignocellulosic

hydrolysates, fed-batch cultivations were carried out in order to simulate the third bioreactor of the three-

stage continuous system and to optimise the cultivation conditions of B. sacchari in xylose.

B. sacchari has been extensively studied as a PHA-accumulating bacteria. More recently, with the

work developed by Raposo et al. (2017), xylitol production has been reported with xylose concentrations

above 30-40 gL-1. Production of xylonic acid has also been observed. These findings revealed alterna-

tive pathways in B. sacchari metabolism when the main carbon source is a pentose. To understand the

metabolic pathways in this bacterial strain several shake flask and bioreactor assays were performed

using xylose as sole carbon source.

In conclusion, this thesis’ work focuses on the xylose metabolism in B. sacchari cultivations, aiming

to find the optimum conditions for P(3HB), xylitol and/or xylonic acid production using this pentose as

substrate.

31

32

4Materials and Methods

33

4.1 Microorganism

Burkholderia sacchari DSM 17165, a strain able to grow, accumulate PHAs, and produce xylitol and

xylonic acid on xylose, was used throughout this work.

4.2 Strain storage and preparation

Cultures of B. sacchari were stored at -80°C in 2 mL cryovials. The stock cultures were prepared

transferring 1500 µL of a previously grown liquid culture in seeding medium (see subsection 4.3.1)

supplemented with 20 gL-1 of xylose, collected in the late exponential growth phase, to sterile cryovials

containing 300 µL of pure sterilized glycerol. This procedure was performed under aseptic conditions

(sterilization under UV light for 15 minutes) in a laminar flow chamber (BIOAIR Instruments aura 2000

MAC 4 NF, Italy) using sterile material.

4.3 Culture media

4.3.1 Seeding medium

The seeding medium (Kim et al., 1994) was prepared by mixing the compounds listed in table 4.1 in 1

L of distilled water, where the pH was adjusted to 6.8 adding the conjugated acid of the phosphate buffer,

KH2PO4. The medium was sterilized by autoclaving at 121°C for 20 minutes. A 100 gL-1 MgSO4·7H2O

solution was prepared and autoclaved separately to avoid the formation of precipitates, and then added

to the medium under sterile conditions.

Table 4.1: Seeding medium composition.

Compound Concentration (gL-1)i Brand name Purity (%)

Na2HPO4·2H2O 4.47 VWR Prolabo 99.9

KH2PO4 1.5 Panreac 98.0-100.5

(NH4)2SO4 1.0 Panreac >99.0

Yeast Extract 1.0 Quilaban −

MgSO4·7H2O solution 2 mLL-1 Panreac 98.0-102.0

Oligo elements solution 1 mLL-1 − −i Unless stated otherwise.

4.3.1.A Oligoelements solution

The oligo elements solution (Kim et al., 1994) was prepared by dissolving the compounds listed in

table 4.2 in distilled water, and then sterilized by autoclaving at 121°C for 20 minutes. This solution was

stored at 4°C.

34

Table 4.2: Oligo elements solution composition.

Compound Concentration (gL-1)i Brand name Purity (%)

FeSO4·7H2O 10 Sigma >99.0

ZnSO4·7H2O 2.25 Sigma >99.0

CuSO4·5H2O 1.00 Panreac >99.0

MnSO4·H2O 0.379 Sigma >99.0

CaCl2·2H2O 2.00 Merck >99.5

Na2B4O7·10H2O 0.23 Merck 99.5-105.0

(NH4)MO7O24·4H2O 0.106 Merck >99.0

HCL 37% 10 mL Fisher Chemical 35i Unless stated otherwise.

4.3.2 Cultivation medium to impose limitation by phosphorus

The bioreactor cultivation medium used to trigger limitation by phosphorus (Kim et al., 1994) was

prepared by mixing the components listed in table 4.3 in distilled water, to attain 1.3 L final working

volume, and adjusting the pH to 7.11 with a 5 M KOH (Panreac) solution. The medium was sterilized

inside the bioreactor at 121°C for 20 minutes. A 100 gL-1 MgSO4·7H2O solution was prepared and

autoclaved separately to avoid the formation of precipitates. The magnesium and a concentrated sugar

solutions were later added to the medium aseptically.

Table 4.3: Phosphorus limited medium composition.

Compound Concentration (gL-1)i Brand name Purity (%)

(NH4)2SO4 4.0 Panreac >99.0

KH2PO4 3.0 Panreac 98.0-100.5

Citric acid·H2O 1.85 Panreac 98.0-102.0

EDTA 40 mg Panreac 98.0

MgSO4·7H2O 12 mLL-1 Panreac 98.0-102.0

Oligo elements solution 10 mLL-1 − −i Unless stated otherwise.

4.3.3 Cultivation medium to impose limitation by nitrogen

The bioreactor cultivation medium used to impose limitation by nitrogen (Kim et al., 1994) was pre-

pared by mixing the components listed in table 4.4 in distilled water, to attain 1.3 L final working volume,

and adjusting the to 7.11 with a 5 M KOH (Panreac) solution. A 100 gL-1 MgSO4·7H2O solution was

prepared and autoclaved separately to avoid the formation of precipitates. The magnesium and a con-

centrated sugar solutions were later added to the medium aseptically.1During autoclaving, the culture medium tends to become more acidic, decreasing below the optimum pH value, 6.8.

35

Table 4.4: Nitrogen limited medium composition.

Compound Concentration (gL-1)1 Brand name Purity (%)

(NH4)2SO4 4.0 Panreac >99.0

KH2PO4 13.3 Panreac 98.0-100.5

Citric acid·H2O 1.85 Pareac 99.5-102.0

EDTA 40 mg Panreac 98.0

MgSO4·7H2O 12 mLL-1 Panreac 98.0-102.0

Oligo elements solution 10 mLL-1 − −1 Unless stated otherwise.

4.4 Carbon sources

All the shake flask and bioreactor assays were carried out using monohydrate dextrose (or D-glucose)

(Dextropam, Portugal), D-sucrose (Fischer Scientific, UK), and D-xylose (Danisco GmbH, Austria). The

sugar solutions were prepared with deionized water and then sterilized by autoclaving at 121°C for 20

minutes. Concerning dextrose (glucose), yield and productivity calculations were performed in terms of

anhydrous glucose.

4.4.1 Feed solution

Unless stated otherwise, the feed for both bioreactors consisted of a 600 gL-1 xylose solution, which

was fed to the stirred-tank reactors (STRs) at a flow rate of ca. 2.7 Lmin-1 (2.6 Lmin-1 for STR 1 and 2.8

Lmin-1 for STR 2.). This solution was prepared with deionized water and then sterilised by autoclaving

at 121°C for 20 minutes. When not in use, the feed solutions were stored at 4°C.

4.5 Culture conditions

4.5.1 Shake flask assays

4.5.1.A Inoculum preparation

Inocula for the shake flasks assays were prepared transferring the content of one cryovial to 500

mL shake flasks containing 94 mL of seeding medium supplemented with 20 gL-1 of glucose (unless

stated otherwise), and incubated at 30°C in an orbital incubator (Infors AG, Switzerland) at 170 rpm for

12 hours (or more, in case of using a different sugar), i.e. until the end of the exponential growth phase.

4.5.1.B Shake flask cultivation

Shake flask assays were performed to determine the growth, substrate consumption and by-products

formation by B. sacchari on xylose. These assays were carried out in 500 mL baffled conical flasks con-

taining 100 mL of liquid phase. The inoculum fraction varied between 5-10% (v/v) in order to obtain

identical initial optical densities (ca. 0.3). Different initial sugar concentrations were used, as well as

36

different carbon sources. These assays were performed in duplicate and the average value was consid-

ered.

4.5.1.C Culture sampling

Culture samples were periodically harvested, four per day on average, in order to analyse biomass,

sugar, polymer and other by-products concentrations. In some cases pH was also measured.

4.5.2 Fed-batch assays

4.5.2.A Inoculum preparation

Inocula for the bioreactor assays were prepared as described in section 4.5.1.A, with a volume of 65

mL (5% v/v of the bioreactor initial working volume).

4.5.2.B Fed-batch cultivation

Fed-batch cultivations were carried out in 2 L STRs (New Brunswick Bioflo 115) operated using the

BioCommand Batch Control software, which enabled control, monitoring and data acquisition. The pH

was controlled at 6.8 with 30% NH4OH or 5 M KOH, depending on the type of limitation used as trigger

for P(3HB) production, namely phosphorus or nitrogen limitation, respectively. The aeration rate used

was 2.6 Lmin-1, unless stated otherwise, and the temperature was set to 30°C. The dissolved oxygen set

point was 20% saturation, unless stated otherwise, and the maximum agitation speed was 1200 rpm.

The inoculum (5% v/v) was prepared as described in section 4.5.1.A. The initial volume of the fed-batch

cultivations was 1.3 L, including all medium components and inoculum.

Feeding was triggered by the increase of the dissolved oxygen (with a correspondent decrease in

the stirring speed), resulting from carbon source exhaustion in the medium. Figure 4.1 presents the

schematics for the experimental set-up of fed-batch cultivations.

Figure 4.1: Experimental set-up for the fed-batch cultivations. Inocula grown on glucose (Gluc), is transferred to the bioreactorwhere it is cultivated on xylose (Xyl) with pH and temperature control, and online DO measurement. Periodic sampling takes placefor biomass, sugars and organic acids, and P(3HB) determination.

37

4.5.2.C Culture sampling

In the fed-batch assays, culture samples were periodically harvested, three per day on average,

in order to determine biomass, sugar, polymer and other by-products concentrations. Phosphate and

ammonia concentrations were also determined when relevant. Typically, a total volume of approximately

17 mL of culture was harvested with a syringe through a non-return valve to maintain aseptia.

4.6 Analytical methods

4.6.1 Optical density measurements

Cellular growth was monitored offline by measuring the optical density (OD) of samples at 600 nm

in a double beam spectrophotometer (Hitachi U-2000), using 3 mL glass cuvettes with an optical path

length of 1 cm. For the OD determination, an aliquot of the culture sample was diluted with deionized

water in order to obtain an absorbance value lower than the threshold (ca. 0.5-0.6). Deionized water

was used as reference.

4.6.2 Cell dry weight determination

The cell dry weight (CDW) was determined by centrifuging at 12000 rpm for 3 minutes (in a Sigma

1-15 P microcentrifuge) 1.2 mL aliquots of culture samples collected in dried and weighted microtubes.

The supernatant was rejected and the pellet washed with deionized water, and then dried at 60°C in

a Memmert oven (Model 400) until constant weight. The CDW was determined dividing the weight

difference after drying the pellets by the collected aliquots volume.

4.6.3 Carbon sources, xylonic acid, xylitol and phosphate determinations

Glucose, sucrose, xylose, xylonic acid, xylitol and phosphate concentrations were determined offline

in a High Performance Liquid Chromatography (HPLC) apparatus (Hitachi LaChrom Elite) equipped with

a Rezex ROA-Organic acid H+ 8% (300 mm × 7.8 mm) column, an autosampler (Hitachi LaChrome

Elite L-2200), an HPLC pump (Hitachi LaChrome Elite L-2130), a Hitachi L-2490 refraction index (RI)

detector for sugars and phosphate and a Hitachi L-2420 UV-Vis detector for organic acids. A column

heater for larger columns (Croco-CIL 100-040-220P, 40 cm × 8 cm × 8 cm, 30-99°C) was connected

externally to the HPLC system. The injection volume was 20 µL and elution was achieved using a 5 mM

solution of H2SO4 as mobile phase. The column was kept at 65°C under a pressure of 26 bar, and the

pump operated at a flow rate of 0.5 mLmin-1.

Due to equippment failure during this thesis’ experimental work, another HPLC had to be used in

order to determine carbon sources and by-products concentrations. This equipment comprised a similar

HPLC apparatus (Hitachi LaChrom Elite), equipped with an Aminex HPX-87H (300 mm × 7.8 mm)

column. The injection volume was 20 µL and elution was achieved using a 5 mM filtered solution of

H2SO4 as mobile phase. The column was kept at 65°C under a pressure of 46 bar, and the pump

38

operated at a flow rate of 0.6 mLmin-1. The two detector modules (RI and UV-vis) were also set in

series.

4.6.3.A Sample preparation

Samples for HPLC analysis were prepared by mixing 300 µL of supernatant aliquots with 300 µL of

a 50 mM solution of H2SO4 in a microtube. After vortexing, these solutions were centrifuged (in a Sigma

1-15 P microcentrifuge) at 12000 rpm for 3 minutes. Samples for injection consisted of 100 µL of the

previous dilution plus 900 µL of the 50 mM H2SO4 solution, giving a final dilution of 1:20.

4.6.3.B Calibration curves

Calibration curves for glucose, xylose, xylonic acid, xylitol and phosphate determinations were ob-

tained for working ranges of 0.5 to 60 gL-1 for glucose, xylose, xylonic acid and phosphate, while a range

of concentrations from 0.1 to 50 gL-1 was chosen for xylitol (see equations A.1 to A.6 in appendix A).

Xylose and xylonic acid showed similar retention times in HPLC runs, which caused an overlap of

both peaks when using the RI detector leading to incorrect concentration values. To avoid this, xylonic

acid concentration was determined using the UV-visible chromatograms (see figure A.2) with an ap-

propriate calibration curve. Using another calibration curve determined for the RI chromatograms (see

figure A.1) the area of xylonic acid in RI chromatograms was computed through the concentration previ-

ously calculated. By subtracting the xylonic acid peak area to the total peak area measured with the RI

detector, it was possible to determine the ”real” xylose peak area and, consequently, the sugar concen-

tration of the sample.

During this experimental work, it was necessary to use a second HPLC apparatus due to technical

problems encountered while using the first, therefore standards were run again in order to determine

new calibration curves. Equations A.7 to A.12 were used to determine the carbon source, xylonic acid,

xylitol and phosphate concentrations.

It is of upmost importance to outline the fact that in the latter HPLC apparatus phosphate and xylitol

exhibit identical retention times. This became an issue in bioreactor cultivations directed towards xylitol

production, as described further in this document.

4.6.4 P(3HB) determination

P(3HB) determination was carried out by Gas Chromatography (GC). Samples prepared as de-

scribed below, in section 4.6.4.A, were analysed in a GC (Agilent Technologies 5890 series II) equipped

with a FID detector and a 7683B injector. The capillary column was a HP-5 from Agilent J&W Scientific,

30 m in length and 0.32 mm of internal diameter. The oven, injector, and detector were kept at constant

temperatures of 60°C, 120°C, and 150°C, respectively. Data acquisition and integration were performed

by a Shimadzu CBM-102 communication Bus Module and Shimadzu GC solution software (version 2.3),

respectively. Peak identification was achieved using as standard 3-methyl hydroxybutyrate (Sigma).

39

4.6.4.A Sample preparation: acidic methanolysis

In order to prepare samples for GC analysis, 1.2 mL aliquots of culture medium were withdrawn from

the culture medium and centrifuged at 12000 rpm for 3 minutes. The pellets were washed with deionized

water and frozen for storage prior to acidic methanolysis. Acidic methanolysis (Cavalheiro et al., 2012)

of the polymer in the cell pellet was carried out by adding to the microtubes 1 mL of an acidic methanol

solution containing (per 100 mL solution): 97 mL of methanol plus 3 mL of H2SO4 96% and 330 µL

of hexanoic acid as the internal standard (IS). The re-suspendend pellets were transferred to Pyrex

hermetic tubes with teflon cases and 1 mL of chloroform was added to each tube. After vortexing for

1 minute, these mixtures were incubated for 5 hours at 100°C in a Memmert GmbH oven (model 200).

After cooling, 1 mL of Na2CO3 was added to the tubes for neutralization, and the samples were vortexed

for 1 min and centrifuged at 4500 rpm for 5 minutes in a Heraeus SEPATECH Labofuge centrifuge

(model 200). 200 µL of the organic phase from each microtube was withdrawn to appropriate vials and

kept at -20°C until GC analysis.

4.6.4.B Calibration curve

The calibration curve for P(3HB) determination was obtained using samples of P(3HB) previously

produced which were subjected to acidic methanolysis, as described before. This curve was determined

for a working range of 0 to 10 gL-1, and is described by equation A.13.

4.6.5 Ammonium quantification

Ammonium concentration was determined offline with an ion-selective electrode (Mettler Toledo In-

Lab 152233000), with a Ag/AgCl reference system in order to confirm the nitrogen limitation in the culture

medium when required.

4.6.5.A Sample preparation

To proceed with the ammonium determination, 15 mL samples of culture broth were harvested pe-

riodically and centrifuged in an Eppendorf 5810 R centrifuge at 12000 rpm during 3 min, at room tem-

perature. The supernatant recovered was then diluted with deionized water to a final volume of 45 mL,

and 5 mL of ISA2 solution were added to maintain the pH constant and minimize interferences with other

ions.

4.6.5.B Calibration curves

Calibration curves with NH4Cl known concentration solutions were performed every time the elec-

trode was used. To assess nitrogen measurement in samples along the entire culture, it was found

necessary to define two working ranges for the ammonium concentrations, a low range from 10-6 M to

10-4 M, and a high range from 10-4 M to 10-1 M. Since this quantification was performed several times,

for different assays, the calibration curves obtained are not included in this section, but can be consulted

in appendix A.2Ionic Strength Adjuster (ISA) is a 0.9 M aluminium sulphate solution (METTLER TOLEDO).

40

4.6.6 Overall yield and productivity calculations

To assess by-products formation during B. sacchari shake flask cultivations, overall product yields

were computed according to equation 4.1. In general, these values were obtained dividing the final

product concentration ([P ]f ) by the concentration of substrate consumed ([S]cons), which is xylose, un-

less stated otherwise. The substrate consumption is determined through a mass balance, computing

the difference between the initial ([S]i) and final ([S]f ) substrate concentrations in the culture broth. It is

important to notice that volume variations due to sampling were neglected in shake flasks assays, i.e.

the biomass concentration is assumed not to be affected.

YP/S (g P/g Scons) =[P ]f

[S]i − [S]f=

[P ]f[S]cons

(4.1)

In bioreactor cultivations, yield calculations were performed using a different approach: instead of

using the final by-product concentration, the mass of product present in each sample harvested was

calculated (assuming that an approximate volume of 17 mL was retrieved at each sampling time), just

as in the culture broth, at the end of the cultivation. In other words, a mass balance was performed at

each sampling time for each product.

As described in equation 4.2, the yields were computed dividing the total by-product mass obtained

in samples (Σ mPs ) plus the remaining amount in the broth (mPf), by the amount of substrate consumed

during cultivation (mScons). Again, the substrate consumption is determined through a mass balance,

computing the difference between the initial amount of substrate (mSi), plus its quantity added in the

feed for fed-batch cultivations (mSfeed), and the final substrate mass in the culture broth (mSf

).

YP/S (g P/g Scons) =Σ mPs

+mPf

mSi+mSfeed

−mSf

=mPf

mScons

(4.2)

Both for shake flask and bioreactor assays, overall productivities were computed dividing the final

product concentration attained ([P ]f ) by the cultivation time (tcult). Similarly, maximum productivities

were computed dividing the maximum product concentration ([P ]max) by the corresponding sampling

time (ts). Equations 4.3 and 4.4 describe these calculations.

ProdP (g P L−1 h−1) =[P ]ftcult

(4.3)

ProdmaxP(g P L−1 h−1) =

[P ]max

ts(4.4)

41

42

5Results and Discussion

43

5.1 Shake flask assays

Fed-batch bioreactor assays based solely on xylose (Raposo et al., 2017) have shown a negligible

P(3HB) production compared to the production on the same carbon source but in shake flasks. With the

aim of understanding the reasons for that fact, several assays were carried out in shake flasks under

different culture conditions.

Shake flask experiments have thus been designed to explore the culture behaviour according to:

i) the composition of cultivation medium, ii) the oxygen availability, and iii) pH control. In addition, the

influence of using inocula grown on different carbon sources on the P(3HB) production based on xylose

has also been tested.

5.1.1 Xylose as sole carbon source for B. sacchari cultivations

To assess the implications in B. sacchari cultivations when using xylose as main carbon source,

a shake flask assay was performed in order to evaluate the growth and production of this strain. The

maximum specific cell growth rate (µmax) of B. sacchari in xylose cultivations was also computed. Figure

5.1 represents the time course of growth, production and substrate consumption obtained in this assay.

Figure 5.1: B. sacchari shake flask cultivation data for growthand production in seeding medium, supplemented with 20 gL-1

of xylose as main carbon source, using an inoculum cultivatedon glucose.

In batch cultivations, the variation of biomass (X), in gL-1, over time is expressed by equation 5.1,

which, when integrated, gives equation 5.2. In these, µ (h-1) is the specific cell growth rate, and X0 (gL-1)

corresponds to the biomass at the beginning of the exponential phase, i.e. at time t0 (h). Since during

the exponential phase cells grow at their maximum growth rate, the plot of experimental data in this time

44

Figure 5.2: Plot of equation 5.2 using experimental data fromthe exponential phase of the shake flask assay represented infigure 5.1.

interval (ln(CDW ) vs. time, as seen in figure 5.2) allowed for µmax (h-1) determination for B. sacchari

shake flask cultivations using xylose as sole carbon source.

dXt

dt= µ×Xt (5.1)

ln(Xt) = ln(X0) + µ (t− t0) (5.2)

From literature, it is known that B. sacchari cultures can achieve a CDW of 6.3 gL-1, with a maximum

production of 3.8 gL-1 of P(3HB), which corresponds to 60% of cell polymer content, when using glucose

as main carbon source (Cesario et al., 2014).

In this experiment, a CDW of ca. 4.2 gL-1 was achieved, with 1.6 gL-1 of P(3HB), corresponding to

39% of cell content. A P(3HB) overall yield of 0.09 g of P(3HB) per g of xylose consumed was obtained,

as well as an overall yield in D-xylonic acid (XylAc) of 0.83 g of XylAc per g of xylose. Additionally,

according to the plot represented in figure 5.2 a µmax of 0.14 h-1 was obtained. This value is slightly

lower than the one reported by Cesario et al. (2014) (µmax of 0.18 h-1) for B. sacchari cultivations on

30 gL-1 of xylose. The lack of more experimental points may be the reason for such difference. In this

report, a µmax of ca. 0.28 h-1 for cultivations on glucose was obtained, which is consistent with the fact

that cell growth on xylose is twice slower than in cultivations on glucose. The yield of P(3HB) on xylose

is not consistent with the one reported in Cesario et al. (2014), where a higher value of 0.24 g P(3HB)

per g xylose was obtained. However, in that study xylonic acid production was not reported as it had

not been yet detected. Due to the overlaping of the xylose and xylonic acid peaks when using the RI

detector, the amount of xylose consumed was miscalculated and so was the yield of polymer on xylose.

In the present study, xylonic acid production was in fact observed as a by-product of xylose metabolism

by B. sacchari , which is in agreement with more recent reports (Raposo et al., 2017).

5.1.2 Influence of the C-source used for inocula growth

In order to evaluate the influence, if any, of the substrate used to grow the inoculum, three shake

flask assays were performed in triplicate, each inoculated with cells cultivated in different sugars, namely:

45

Figure 5.3: Experimental set-up to study of the influence of the sugar used to grow the inoculum: glucose (Gluc), xylose (Xyl)and sucrose (Suc), on the growth and production performance of cultures grown on 20 gL-1 of xylose. Periodic sampling for pH,biomass, sugars and organic acids, and P(3HB) determination.

Table 5.1: Overall yields of xylonic acid (XylAc) and P(3HB) in shake flask cultivations using 20 gL-1 of xylose as sole carbonsource, and inocula grown in three different sugars: glucose, xylose and sucrose.

C-sourceYXylAc/Xyl cons YP(3HB)/Xyl cons %P(3HB)

(g XylAc/g Xyl) (g P(3HB)/g Xyl) (g P(3HB)/g CDW)

Glucose 0.54 0.086 39.4

Xylose 0.54 0.084 39.8

Sucrose 0.68 0.090 43.6

glucose (Gluc), D-xylose (Xyl) and sucrose (Suc). A schematic of the experimental set-up is represented

in figure 5.3.

Figures 5.4 to 5.6 represent the time course of production and substrate consumption in shake flask

cultivations, and table 5.1 summarises the results in terms of product overall yields and polymer content.

As shown in figures 5.4 to 5.6, the residual biomass (Xr) values attained are similar for all three

cultivations (around 2 gL-1). Inocula grown on glucose and xylose feature similar substrate (xylose)

consumption during both cultivations, while the culture with the inoculum grown on sucrose led to higher

consumption of xylose.

Despite the absence of glucose in the data represented above, a residual amount of this sugar was

present in the assays due to the carryover of glucose from the inoculum. This explains the fact that xylose

is not being consumed from the very beginning. Since it was a negligible amount of sugar (less than 1

gL-1), its presence was not taken into account in any yield calculations. Since sucrose is a disaccharide

46

Figure 5.4: B. sacchari shake flask cultivation data for growthand production in SM, supplemented with 20 gL-1 of xylose asmain carbon source, and an inoculum grown on glucose.

Figure 5.5: B. sacchari shake flask cultivation data for growthand production in SM, supplemented with 20 gL-1 of xylose asmain carbon source, and an inoculum grown on xylose.

Figure 5.6: B. sacchari shake flask cultivation data for growthand production in SM, supplemented with 20 gL-1 of xylose asmain carbon source, using an inoculum grown on sucrose.

47

composed of glucose and fructose units, it wasn’t possible to determine this sugar concentration. As

such, sucrose presence in shake flask cultivations was also assumed as negligible.

Table 5.1 shows identical D-xylonic acid (XylAc) yields in xylose cultivations using inocula grown on

glucose or xylose, and higher yield for the inoculum grown on sucrose. The same results are again ob-

tained for P(3HB) yields and %P(3HB). Also, the polymer content, around 40% of the CDW, is consistent

with the assay described in section 5.1.1 and also with the report of Cesario et al. (2014).

The results obtained suggest that in B. sacchari xylose-cultivations there are no significant differ-

ences in terms of residual biomass and product overall yields using inocula grown on glucose or xylose.

However, when sucrose grown inocula was used, higher cell growth and products concentrations were

attained, due to an increased quantity of substrate consumed. Despite these conclusions, in order to

compare the results of this study with previous studies, it was chosen to proceed the experimental work

with glucose grown inocula.

Moreover, B. sacchari cell growth on glucose, xylose and sucrose was monitored during inocula

preparation. It was interesting to notice that B. sacchari cultures on xylose achieve the stationary phase

only after ca. 24h, i.e. much later than those on glucose and sucrose. This is in fact supported by

the lower µmax determined experimentally for cultivations on xylose (see section 5.1.1), and also by

literature (Cesario et al., 2014)).

5.1.3 Influence of YE presence in seeding medium

As mentioned previously, another significant difference between shake flask and bioreactor media is

the fact that the first is a complex medium whereas the second is a defined medium.

The rate of growth and the activity of metabolic processes may be strongly affected by the type and

ratio of nutrients provided to the culture, hence affecting cell mass and specific product yields. These

nutrients are often supplied from raw materials that are complex, ill-defined mixtures of natural origin and

subject to significant variation, such as yeast extract (YE). YE, produced by the hydrolysis and spray-

drying of baker’s or brewer’s yeast, is a powder consisting of protein, free amino nitrogen, vitamins, and

minerals (Kasprow et al., 1998).

Aiming to study the effect of YE presence in the culture medium, an assay in duplicate was carried

out to compare cell growth and product yields of B. sacchari cultivations on xylose. A schematic of the

experimental set-up is represented in figure 5.7.

Figures 5.8 and 5.9 represent the time course of production and substrate consumption, and table

5.2 summarises the results in terms of product overall yields and polymer content.

As stated in the beginning of this section, cell growth is highly dependent on the nutrients ratio

present in the culture medium. If the amount of essential nutrients is too small for the amount of carbon

source available, as it happens in the culture without YE, the sugar consumption will be lower, leading

to a lower cell growth and a lower polymer production.

The results displayed in table 5.2, indicate that the xylonic acid yield is similar in both cases, but

around 10% higher in the cultivation carried without YE, suggesting that the production of this organic

acid is not directly influenced by the nutrients provided through this compound. Additionally, this be-

48

Figure 5.7: Experimental set-up for the study of the influence of yeast extract (YE) presence in the shake flask cultivation seedingmedium. Inoculum grown on glucose (Gluc) cultivated in complete seeding medium (SM) and in seeding medium without YE(SM-YE), supplemented with 20 gL-1 of xylose (Xyl). Periodic sampling for pH, biomass, sugars and organic acids, and P(3HB)determination.

Figure 5.8: B. sacchari glucose grown inoculum cultivated inshake flask, in complete SM supplemented with 20 gL-1 of xyloseas main carbon source.

Figure 5.9: B. sacchari glucose grown inoculum cultivated inshake flask with 20 gL-1 of xylose as main carbon source, inseeding medium without YE.

Table 5.2: Overall yields of xylonic acid (XylAc) and P(3HB) in shake flask cultivations using seeding medium with and withoutyeast extract (YE).

Cultivation YXylAc/Xyl cons YP(3HB)/Xyl cons %P(3HB)

medium (g XylAc/g Xyl) (g P(3HB)/g Xyl) (g P(3HB)/g P(3HB))

SM 0.61 0.081 38.5

SM-YE 0.72 0.061 35.0

49

haviour indicates that the cells seem to conduct the xylose preferably towards xylonic acid production

pathway instead of P(3HB) formation pathway.

It is important to notice the small differences in terms of residual biomass and product yields between

the cultivations represented in figures 5.4 and 5.8. These cultivations were carried under the same

growth conditions, however, the results obtained for the latter assay show higher XylAc yield and lower

polymer yield. One plausible explanation could be experimental error. The product yields attained for

the cultivation of glucose grown inoculum in complete seeding medium (see table 5.2) are similar to the

product yields attained for the sucrose grown inoculum cultivation (see table 5.1). This supports the idea

of no influence of the sugar used for inoculum cultivations.

5.1.4 Influence of inhibitory substrate concentrations

Previous studies (Raposo et al., 2017) revealed D-xylitol (Xylt) production, both in shake flask and

bioreactor cultivations, when high concentrations of xylose were present, i.e. above 30-40 gL-1.

In order to obtain xylitol production in B. sacchari cultivations, duplicate assays were carried with a

glucose grown inoculum cultivated in seeding media supplemented with 50 gL-1 of xylose. A schematic

of the experimental set-up is represented in figure 5.10. Figure 5.11 represents the time course of

production and substrate consumption, and table 5.3 summarises the results in terms of product overall

yields and polymer content.

Figure 5.10: Experimental set-up for the study of xylitol production in shake flask cultivations. Inoculum grown on glucosecultivated in seeding medium (SM) supplemented with 50 gL-1 of xylose (Xyl). Periodic sampling for pH, biomass, sugars andorganic acids, and P(3HB) determination.

Table 5.3: Overall yields of xylonic acid (XylAc), xylitol (Xylt) and P(3HB) in shake flask cultivations with inhibitory concentrationsof xylose (50 gL-1).

YXylAc/Xyl cons YXylt/Xyl cons YP(3HB)/Xyl cons %P(3HB)

(g XylAc/g Xyl) (g Xylt/g Xyl) (g P(3HB)/g Xyl) (g P(3HB)/g CDW)

0.74 0.037 0.042 28.6

Similar values for residual biomass were obtained for both cultivations with 20 gL-1 and 50 gL-1 (ca.

2.5 gL-1), as represented in figures 5.1 and 5.11, respectively. However, figure 5.11 shows that only half

of the xylose initially present in the culture medium was consumed, probably due to cell growth having

cessed as a consequence of the medium becoming too acidic.

A comparison between the results described in section 5.1.1 and table 5.3 reveals that lower polymer

yields were obtained when the xylose concentration in the culture medium was inhibitory, hence the

50

Figure 5.11: Glucose grown inoculum cultivated in seedingmedium, supplemented with high concentrations of substrate (50gL-1 of xylose).

polymer content was lower than 30%. The same result was obtained for xylonic acid: a lower yield was

found in the presence of 50 gL-1 of xylose. These results demonstrate that less substrate was directed

towards the formation of these two products, suggesting the presence of a third metabolite.

As expected, it was possible to confirm xylitol production in shake flask cultivations of B. sacchari

when using inhibitory concentrations of xylose as main carbon source. Thus, the substrate was con-

ducted towards xylonic acid, polymer and xylitol production, three by-products of the xylose metabolism.

5.1.5 Influence of pH

The enzymatic activity is of microbial cells is intrinsically related with the medium pH. Bacterial cells

are capable of maintaining their intracellular pH even when subjected to significant exterior pH varia-

tions. This happens at the expense of maintenance energy, so as to keep the proton gradient constant

through the cell membrane. However, this capacity decreases when the extracellular pH drops due to

organic acids formation. These acids penetrate cells by passive diffusion, where they meet a higher

intracellular pH and dissociate. High organic acids concentrations force cells to pump protons to the

exterior, leading to an elevated energy outlay (da Fonseca and Teixeira, 2007). Medium buffering is

almost always insufficient to maintain the culture pH between acceptable values, hence the addition of

acids or bases to control the pH is essential.

To assess pH variations during culture growth, shake flask assays were performed. Foremost, the

goal was to control the culture medium pH with small additions of a base solution with known concen-

tration (1 M KOH). However, due to the presence of the phosphate buffer in the seeding medium, it was

51

not possible to maintain the pH at the desired value and relevant results were not obtained. As such,

this study was later carried out in bioreactor (see section 5.2.1), where pH control is effective.

Despite this, it was possible to observe how the pH varies during culture growth in shake flasks. As

stated in section 4.3.1, the shake flask medium was adjusted to 6.8. Since one of the main products of

B. sacchari metabolism in xylose is an organic acid, its accumulation in the broth during the cultivation

leads to a decrease of the pH, and a too acidic medium for the bacteria to grow on.

Figure 5.12 represents the pH variation with OD of the culture. It shows the increasing acidity in the

medium culture during B. sacchari cultivation in shake flask, to a final pH of 3.2. This assay indicates

that, in shake flask cultivations in 20 gL-1 of xylose, this bacterium grows to a maximum OD of 13 (cor-

responding to a CDW of 4 gL-1), for approximately 42h, when the accumulation of metabolites becomes

inhibitory. This pH profile was also observed in previous assays, as shown for example in figure 5.1.

Figure 5.12: B. sacchari pH profile during shake flask cultivationin seeding medium, supplemented with 20 gL-1 of xylose.

5.2 Bioreactor assays

5.2.1 Influence of pH in B. sacchari batch cultivations

To study the pH influence on culture growth, bioreactor cultivations were performed in batch mode

using the same cultivation medium as in shake flask assays, i.e. seeding medium (SM). Two assays took

place, one with automatic pH control and another without it. In the first case the set/point was 6.8, and

the base added was a 5 M KOH solution and not a NH4OH solution, so that the polymer accumulation

could be triggered by nitrogen limitation.

Figures 5.13 and 5.14 represent the time course of growth and production, as well as the acquisition

data from the batch cultivations A and B. Table 5.4 summarises the results obtained in terms of product

overall yields and productivities, and P(3HB) content.

Figure 5.13 shows that all the xylose (Xyl) available in the medium was consumed when the pH

of the culture was controlled, whereas in the cultivation without pH control, represented in figure 5.14,

approximately around half of the initial amount of carbon source was left unused. In this case the

xylose metabolism ceased after circa 35h, hence the cell growth also stalled, due to the acidity in the

culture medium, a behaviour also observed in shake flask assays (see figure 5.1). Figure 5.13 also

shows that, after a certain time of cultivation, when growth and polymer accumulation stops, xylose is

directed towards xylonic acid (XylAc) and xylitol (Xylt) production only. However, these two metabolites

52

Figure 5.13: Data obtained in B. sacchari batch cultivation A,using seeding medium supplemented with 50 gL-1 of xylose, withpH controlled at 6.8 and 20% of DO. Equations A.1 to A.6 wereused as calibrations curves.

Figure 5.14: Data obtained in B. sacchari batch cultivation B,using seeding medium supplemented with 50 gL-1 of xylose, with20% of DO and without pH control. Equations A.1 to A.6 wereused as calibrations curves.

Table 5.4: Overall yields and productivities of xylonic acid (XylAc), xylitol (Xylt) and P(3HB) in bioreactor batch cultivations, usingseeding medium supplemented with 50 gL-1 of xylose, with (A) and without (B) pH control, and 20% saturation of DO in themedium.

Cultivation XylAc Xylt P(3HB)%P(3HB)

(g P(3HB)/g CDW)

A: with pH control

YP/S (g/g Xyl) 0.95 0.056 0.039

ProdP/S (g L-1 h-1) 0.45 0.027 0.018 39

Prodmax P/S (g L-1 h-1) 0.46 0.031 0.018

B: without pH control

YP/S (g/g Xyl) 0.54 0.027 0.029

ProdP/S (g L-1 h-1) 0.18 0.009 0.009 35

Prodmax P/S (g L-1 h-1) 0.22 0.010 0.047

53

start being formed even before polymer accumulation starts. The shortage of some component in the

exhausted culture medium may be the cause for the cessation of P(3HB) production stopping and thus

xylose is channeled only to xylitol and xylonic acid production.

Theoretically, the more substrate cells consume, the higher the cell or product concentration achieved.

In the batch assay with pH control at 6.8, is possible to confirm the higher cell growth (ca. 3.6 gL-1 of

residual biomass) and production levels (approximately 60 gL-1 of xylonic acid and ca. 3.6 gL-1 of xylitol)

compared to the cultivations without pH control. The results concerning P(3HB) production are higher

when pH control takes place probably since more substrate is consumed and channeled into polymer

formation. In cultivations without pH control, the metabolism tends to stagnate at pH 3.2, the cell growth

ceases and so does the metabolite production. These results establish the necessity of pH control for

microbial cultivations.

Note that the pH fluctuation represented in figure 5.14 is identical to the pH profile obtained for

B. sacchari in shake flask (see figures 5.1 and 5.11). Also, from the results described in table 5.4, the

cell content of P(3HB) is similar to the values attained for shake flask assays (see table 5.1). Another

interesting aspect is xylitol production, which occurs when high concentrations of xylose are present in

the culture medium, as reported in the literature (Raposo et al., 2017) and confirmed in the correspond-

ing shake flask assay (see section 5.1.4).

At fixed air flow, the agitation in bioreactor cultivations changes automatically to maintain the dissolved

oxygen (DO) set-point (%DO sat.). The concentration of the DO in the culture medium is a translation of

the cells metabolism. In other words, when cells are metabolising the sugar, oxygen is being consumed,

hence the agitation increases to maintain the DO set-point, i.e. at 20% saturation. If there is no substrate

left for consumption, or cell growth just ceased due to metabolite accumulation in the broth, the %DO

increases towards saturation and agitation levels remain at the lower threshold.

Data regarding agitation and DO in the culture medium were collected during the time course of both

cultivations. Observing figures 5.13 and 5.14, the most predominant difference is the DO curve in the

cultivation without pH control. Once the metabolism ceases, after 32h of cultivation, the % DO increases

to saturation levels (around 95%) and the agitation remains at the lower threshold. On the contrary,

in the cultivation with pH control, the cells continue metabolizing the remaining substrate, and the DO

presents values around 30% saturation, without the necessity of increasing the agitation.

5.2.2 Influence of the limiting nutrient in the medium used for fed-batch cultiva-tions

As described before, bioreactor cultivation medium is a defined medium (see section 4.3), as op-

posed to the seeding medium used in shake flasks, which is a complex medium due to the presence of

yeast extract. Since P(3HB) synthesis occurs under limitation of an essential nutrient and in the pres-

ence of excess carbon source (Sudesh et al., 2000), two types of bioreactor media were used throughout

this work to impose nutrient limitation: phosphorus and nitrogen limiting media. Another difference be-

tween seeding and the bioreactor media tested in previous studies (Raposo et al., 2017) is that the first

54

is limited by nitrogen while the latter is limited by phosphorus. Recent research on P(3HB) production

using B. sacchari in bioreactor cultivations has been carried out in media with P-limitation (Silva et al.,

2004; Cesario et al., 2014) since (Ryu et al., 1997) first reported that higher polymer accumulation was

attained on Ralstonia eutropha cultivations based on glucose.

In order to evaluate the influence of the limiting nutrient in the medium used for B. sacchari cultiva-

tions in xylose, fed-batch assays were performed with different media compositions (see section 4.3)

and the results obtained for growth and production compared. Figures 5.15 and 5.16 represent the time

course of growth and metabolite production, as well as data obtained in the fed-batch cultivations C

and D. Table 5.5 summarises the results in terms of product overall yields and productivities, as well as

P(3HB) content.

Figure 5.15: Data obtained in B. sacchari fed-batch cultivationC, using P limited medium supplemented with 30 gL-1 of xyloseas main carbon source, at pH 6.8 and 20% of DO. A 600 gL-1

solution of xylose was used as feed. Equations A.7 to A.12 wereused as calibration curves.

Figure 5.16: Data obtained in B. sacchari fed-batch cultivationD, using N limited medium supplemented with 30 gL-1 of xyloseas main carbon source, at pH 6.8 and 20% of DO. A 600 gL-1

solution of xylose was used as feed. Equations A.1 to A.6, andA.14 were used as calibrations curves.

55

Table 5.5: Overall yields and productivities of xylonic acid (XylAc), xylitol (Xylt) and P(3HB) in fed-batch cultivations C and D,carried out in an essential nutrient limiting medium supplemented with 30 gL-1 of xylose, at pH 6.8 and 20% of DO in the medium.

Cultivation XylAc Xylt P(3HB)%P(3HB)

(g P(3HB)/g CDW)

C: P-limited

YP/S (g/g Xyl) 0.65 - 0.000

ProdP/S (g L-1 h-1) 0.54 - 0.00 0

Prodmax P/S (g L-1 h-1) 0.92 - 0.016

D: N-limited

YP/S (g/g Xyl) 1.07 - 0.041

ProdP/S (g L-1 h-1) 0.73 - 0.028 35

Prodmax P/S (g L-1 h-1) 0.73 - 0.068

First of all, it is important to point out the alkaline pH shift in the bioreactor assays’ data graphs repre-

sented above within the first 12-24 hours. This could be explained by the citrate present in the medium

that is consumed in the first place (Cesario et al., 2014), and/or by the alkaline shift caused by D-xylose

presence in the culture (Jeffries, 1983; Lam et al., 1980). The latter may support the assumption that

D-xylose uptake in B. sacchari is indeed driven by a chemiosmotic system, as seen for Escherichia

coli (E. coli). Note that the batch assays’ data also exhibit this pH change (see figures 5.13 and 5.14),

even with the presence of buffer.

Figure 5.15 shows no significant polymer formation, suggesting that all the xylose (Xyl) provided was

consumed and used mainly for cell growth (ca. 35 gL-1 of Xr) and xylonic acid production (reaching a

maximum value of 72 gL-1). Phosphorus limitation during cultivation is confirmed, however it did not

promote a significant amount of polymer formation.

On the contrary, figure 5.16 exhibits higher P(3HB) and xylonic acid productivities attained under

nitrogen limitation, as well as polymer content, despite the lower cell growth compared to cultivation C

(around 7.5 gL-1 of residual biomass, less than half of the value attained for medium limited by phos-

phorus). A decrease in the ammonium concentration is also observed in cultivation D, suggesting a

tendency towards nitrogen limitation. With regard to the nutrient limitation, it is important to keep in mind

that the limiting factor could not be simply the nitrogen availability, but instead the ratio carbon-nitrogen.

For instance, if the amount of nitrogen present is too small for the quantity of carbon source available for

cell growth, we should consider that the culture is facing limiting conditions.

Figure 5.16 shows that the pH in cultivation D was measured low in the first 8 hours of culture. Since

the cultivation medium is settled at pH 7.1 prior to sterilization (as described in sections 4.3.3 and 4.3.3),

it is very unlikely that the pH droped so abrutly during this procedure. One explanation could be an

improper assembly of the electrode cable or poor pH calibration. Nevertheless, the culture growth did

not seem compromised and plausible results were in fact obtained. It is also important to highlight that

there were difficulties when using the ammonium electrode to assess N-limitation during the cultivation

once the obtained ammonia concentration values in the culture were erratic. One explanation could be

56

interference from the other components of the broth, such as metabolites, impairing the accuracy of the

equipment.

In short, the results discussed above allow to conclude that cultivations in media limited by nitrogen

lead to higher P(3HB) accumulation when using xylose as main carbon source, unlike what happens

in glucose-cultivations. In fact, Cesario et al. (2014) reported that the %P(3HB) content on xylose was

44.4% compared to glucose, which was determined as 60.3%. On the other hand, cultivations carried

out in phosphorus limited medium led to higher cell growth when compared to their N limited counter-

parts.

One of the objectives for this work assignment is the design of a process approach for P(3HB)

production by B. sacchari on xylose. Given the results obtained so far, a strategy that could be worth

testing would be carrying a first cultivation stage using phosphorus limited medium in order to reach high

values of cell growth, despite the no (significant) polymer formation. Once CDW values above 30 gL-1

had been reached, a second stage of the culture would take place, where the addition of a P-source

(through a feed solution or by changing the base to KOH) would switch the limiting nutrient into nitrogen,

thus promoting P(3HB) production.

5.2.3 Influence of the dissolved oxygen availability in the culture broth

In aerated bioreactors, molecular oxygen is the final electron acceptor in cell metabolism, enabling

fast substrate conversion. Given the high reaction volumes, and the high costs associated with aeration,

bioreactors are often operated with dissolved oxygen (DO) percentages just above the critical concen-

tration, where double limitation from the substrate as well as from the oxygen may occur (da Fonseca

and Teixeira, 2007).

To assess the influence of oxygen availability in B. sacchari cultivations, assays were performed with

a lower value for the DO set-point. Typically, cultivations were carried with an aeration of 2 vvm, and 20%

of DO saturation maintained by the agitation speed (the higher the agitation, the higher the superficial

oxygen transfer area, ergo the higher the rate of oxygen transfer into the medium). In order to decrease

the amount of oxygen available, the DO set-point was set to 1% saturation. To attain this value, the

air flow was reduced to ca. 1.0 Lmin-1, corresponding to 0.8 vvm. It is important to keep in mind that

lowering the amount of dissolved oxygen in the medium does not necessarily mean that the culture is

limited by oxygen itself, but again by high values for the ratio C/O2. With the objective of assessing

the effects of reducing the oxygen supply during cultivation, fed-batch cultivations were carried out in

duplicate using media limited by phosphorous and nitrogen.

The results obtained in the time course for growth and production, as well as online acquired data

in the P-limited fed-batch cultivations E and F are shown in figures 5.17 and 5.18, and figures 5.19

and 5.20 represent the results achieved under N-limitation, corresponding to cultivations G and H, all

with low DO set-point. Table 5.6 summarises the results obtained in terms of product overall yields and

productivities, as well as P(3HB) content.

57

Figure 5.17: Data obtained in B. sacchari fed-batch cultivationE, using P limiting medium supplemented with 30 gL-1 of xyloseas main carbon source, at pH 6.8 and 1% of DO. A 600 gL-1

solution of xylose was used as feed. Equations A.1 to A.6 wereused as calibrations curves.

Figure 5.18: Data obtained in B. sacchari fed-batch cultivationF, using P limiting medium supplemented with 30 gL-1 of xyloseas main carbon source, at pH 6.8 and 1% of DO. A 600 gL-1

solution of xylose was used as feed. Equations A.7 to A.12 wereused as calibrations curves.

58

Figure 5.19: Data obtained in B. sacchari fed-batch cultivationG, using N limiting medium supplemented with 30 gL-1 of xyloseas main carbon source, at pH 6.8 and 1% of DO. Manual pulsesof a 600 gL-1 solution of xylose were added during cultivation asfeed. Equations A.7 to A.12, and A.15 were used as calibrationscurves.

Figure 5.20: Data obtained in B. sacchari fed-batch cultivationH, using N limiting medium supplemented with 30 gL-1 of xyloseas main carbon source, at pH 6.8 and a set point for DO of 1%. Atotal volume of 316 mL of a 600 gL-1 xylose solution was used asfeed. Equations A.7 to A.12, and A.16 were used as calibrationscurves.

59

Table 5.6: Overall yields and productivities of xylonic acid (XylAc), xylitol (Xylt) and P(3HB) in fed-batch cultivations using phos-phorus (cultivations E and F) and nitrogen (cultivations G and H) limited medium supplemented with 30 gL-1 of xylose, at pH 6.8and 1% saturation of DO.

Cultivation XylAc Xylt P(3HB)%P(3HB)

(g P(3HB)/g CDW)

E: P-limited

YP/S (g/g Xyl) 0.17 0.007 0.070

ProdP/S (g L-1 h-1) 0.14 0.005 0.056 26

Prodmax P/S (g L-1 h-1) 0.14 0.005 0.139

F: P-limited

YP/S (g/g Xyl) 0.17 - 0.067

ProdP/S (g L-1 h-1) 0.15 - 0.057 20

Prodmax P/S (g L-1 h-1) 0.23 - 0.058

G: N-limited

YP/S (g/g Xyl) - - 0.120

ProdP/S (g L-1 h-1) - - 0.037 36

Prodmax P/S (g L-1 h-1) - - 0.039

H: N-limited

YP/S (g/g Xyl) 0.29 - 0.083

ProdP/S (g L-1 h-1) 0.18 - 0.047 30

Prodmax P/S (g L-1 h-1) 0.18 - 0.074

As figures 5.17 and 5.18 show, in the P-limited cultivations there is polymer formation for DO con-

centrations around 1%, reaching up to 0.07 g of P(3HB) per g of xylose consumed in cultivation E, and

similar values for CDW are attained. This suggests that a possible double limitation by phosphorus and

oxygen took place and led to P(3HB) production. To avoid a possible limitation by carbon source, man-

ual pulses were added to the culture E which caused an increase in the xylose concentration levels (up

to 40 gL-1) and, as demonstrated previously throughout this work, high xylose concentrations triggered

xylitol production. It is important to notice that it was not possible to verify phosphorus limitation during

cultivation F due to peak overlap in the HPLC chromatograms. Another interesting aspect is the xylonic

acid production, which starts later compared to cultivations with similar limitation but carried out at 20%

saturation of DO in the medium, when xylose concentrations are close to zero. From table 5.6 is possi-

ble to observe similar P(3HB) yields for both assays (ca. 0.07 g P(3HB) per g of xylose consumed), but

lower polymer content in cultivation F.

As discussed previously, culture media limited by nitrogen favoured P(3HB) formation, with a yield of

ca. 0.04 g of P(3HB) per g of xylose consumed for cultivation D. Additionally, with the results represented

in figures 5.19 and 5.20, by decreasing the DO concentration to 1-3% saturation, the polymer production

attained increased up to 0.12 g of P(3HB) per g of xylose consumed, for cultivation G. Therefore, the

P(3HB) yield apparently increases under double limitation, despite the lower cellular growth . As figure

5.19 shows, feeding through the usual strategy was difficult once the air supplied was enough to satisfy

the cellular metabolism, and increasing agitation was not necessary. In consequence, a very small

amount of feed was added to the culture manually, to avoid depletion of the carbon source. Comparing

60

both cultivations G and H, the most evident differences are the cell growth, oxygen and agitation curves.

Due to a possible malfunction of the oxygen electrode or poor calibration, negative values for the DO in

the medium were recorded in cultivation H, which are not physically possible. These triggered the stirrer

speed to unnecessary high values, and thus to frequent feed pulses, hence the xylose accumulation

in the broth in the end of cultivation. However, not enough time passed to observe xylitol production.

Also, the higher oxygen availability in the medium may be responsible for the higher cellular growth.

In addition, P(3HB) production is lower in cultivation H, and lower yield and productivity values were

attained. It is worthy of notice that the ammonia concentration appears to be constant during the time

course of both cultivations. Since this assay was carried under oxygen limitation, cell growth is much

slower and hence more time is needed to achieve the N-limiting conditions.

One final observation regarding the cultivation H is the inconsistency of CDW values, which is likely

due to experimental error during sampling. Therefore, an average value is presented in table 5.6 for the

polymer content.

From the results summarised in table 5.6 in terms of yield and productivity, these indicate that,

among all growth conditions studied, the best approach towards P(3HB) production by B. sacchari , so

far, seems to be a combination of N-limited medium and 1% of oxygen saturation. Therefore, aiming at

an improvement of polymer production based on xylose, a two stage system should take place: a first

stage using P-limiting conditions to achieve high cellular growth, followed by a second cultivation stage

using N-limitation and low values of DO (less than 5%).

5.2.4 Cultivations towards xylitol production: influence of inhibitory xylose con-centrations

To study the parameters that promote xylitol production in B. sacchari cultivations on xylose, du-

plicate assays were performed using medium limited by phosphorus and 20% saturation of DO, since

these were the conditions that led to higher cell growth, with no polymer formation thus increasing the

chances of conducting the substrate consumed towards xylitol formation and secretion.

Figures 5.21 and 5.22 represent the time course of growth and production of cultivations I and J,

respectively, and table 5.7 summarises the results in terms of product overall yields and productivities,

as well as P(3HB) content.

Observing both figures 5.21 and 5.22, these show an identical behaviour in terms of cell growth,

polymer formation and oxygen and agitation profiles. Also, culture medium limitation by phosphorus

is confirmed in both cultivations. However, the feeding strategy used for xylose accumulation in the

broth was not accomplished, thus leading to no xylitol production in any of the assays. The lack of

xylose monitoring during the cultivation (due to HPLC technical problems) was the main cause for the

triggering of feed not having been adjusted correctly. Increasing the pulses’ volume or trigger the xylose

pulses more frequently would have allowed for substrate accumulation in the culture medium, potentially

creating the ”ideal” cultivation conditions for xylitol formation.

Despite these assays, in previous cultivations xylose did accumulate in the broth (see cultivation E

61

Figure 5.21: Data obtained in B. sacchari fed-batch cultivationI, using P-limited cultivation medium supplemented with 30 gL-1

of xylose as main carbon source, at pH 6.8 and 20% of DO. Atotal volume of 198 mL of a 600 gL-1 xylose solution was usedas feed. Equations A.7 to A.12 were used as calibrations curves.

Figure 5.22: Data obtained in B. sacchari fed-batch cultivationJ, using P-limited cultivation medium supplemented with 30 gL-1

of xylose as main carbon source, at pH 6.8 and 20% of DO. Atotal volume of 236 mL of a 600 gL-1 xylose solution was usedas feed. Equations A.7 to A.12 were used as calibrations curves.

Table 5.7: Overall yields and productivities of xylonic acid (XylAc), xylitol (Xylt) and P(3HB) in the bioreactor fed-batch cultivationsusing a phosphorus limited medium, supplemented with 30 gL-1 of xylose, at pH 6.8 and 20% saturation of DO.

Cultivation XylAc Xylt P(3HB)% P(3HB)

(g P(3HB)/g CDW)

I: P-limited

YP/S (g/g Xyl) 0.151 - 0.008

ProdP/S (g L-1 h-1) 0.187 - 0.014 3

Prodmax P/S (g L-1 h-1) 0.195 - 0.014

J: P-limited

YP/S (g/g Xyl) 0.000 - 0.000

ProdP/S (g L-1 h-1) 0.000 - 0.000 0

Prodmax P/S (g L-1 h-1) 0.06 - 0.021

62

shown in figure 5.17), however not intentionally, conducting to xylitol production. The aim of the assays

described above was to achieve high xylose concentrations in the broth and determine the conditions

to enhance xylitol production in B. sacchari cultivations. Despite the efforts made, the influence of high

xylose concentrations was not assessed properly and relevant conclusions weren’t withdrawn from the

assays performed in this matter.

On the other hand, once again it was noted that cultivations using P limiting medium with 20% of DO

did not led to significant polymer formation, confirming the conclusions drawn from the assays performed

earlier.

One interesting aspect worth noticing, is the fact that in the cultivation H, represented in figure 5.22,

cell growth appears to have two exponential phases, given the increasing curve for residual biomass.

This was the highest cell growth achieved within all the work described in this document, ca. 47 gL-1.

Also, the xylonic acid formed appears to be converted given the decreasing concentration values towards

zero, hence the null yield. The increasing pH in the culture medium is an important indicator in terms of

organic acids production and following consumption as well.

Finally, with regard to xylonic acid production, the results obtained throughout this work, may suggest

that when the xylose concentration in the culture medium is limitant, i.e. close to zero, xylonic acid

production takes place (see figures 5.15 to 5.18) whereas xylose concentrations above 30 gL-1 channel

the substrate towards the xylitol formation pathway (see figures 5.13, 5.14 and 5.17). Therefore, an

appropriate cultivation strategy to enhance xylonic acid production involves the control of the xylose

concentration in the broth.

5.2.5 Three-stage continuous bioreactors series: third bioreactor optimisationapproach

As mentioned previously, a three-stage stirred-tank reactor system operating in continuous and aim-

ing at P(3HB) production was run by Henriques (2015) based on a mixture glucose and xylose mimicking

a hemicellulosic hydrlysate. However, significant polymer production was not attained in the third biore-

actor of the series. To find a possible solution for this set back, P(3HB) production was assessed is

a fed-batch cultivation using similar feed composition and cultivation conditions (P-limiting medium and

20% of oxygen saturation). The feed solution was composed by a mixture of xylose and glucose with

a ratio of 8:1. Figure 5.23 represents the time course of growth and production, as well as online data

from the batch cultivations for the assay K.

The assays described in section 5.2.3 demonstrated that a reduced availability of oxygen in cultiva-

tions using P-limiting media leads to polymer production. In this context, the assay L, simulating the third

bioreactor of the continuous series, was also performed at 1% DO, aiming to attain higher P(3HB) yield

and productivites. The results of this cultivation are shown in figure 5.24.

Table 5.8 summarises the results in terms of product overall yields and productivities, as well as

P(3HB) content, for both cultivations described above.

Likewise to the third stage of the continuous series, there was no P(3HB) formation under the sim-

ulated cultivation conditions (P limiting medium with 20% of DO). This suggests that the presence of

63

Figure 5.23: Data obtained in B. sacchari fed-batch cultivationK, using P limiting medium supplemented with 30 gL-1 of xylose,at pH 6.8 and 20% of DO. A mixed solution of 600 gL-1 of xyloseand 75 gL-1 of glucose was used as feed. Equations A.7 to A.12were used as calibrations curves.

Figure 5.24: Data obtained in B. sacchari fed-batch cultivationL, using P-limited medium supplemented with 30 gL-1 of xylose,at pH 6.8 and 1% of DO. A total volume of 26 mL of a mixedsolution of 600 gL-1 of xylose and 75 gL-1 of glucose was usedas feed. Equations A.7 to A.12 were used as calibrations curves.

Table 5.8: Overall yields and productivities of xylonic acid (XylAc) and P(3HB) in the third bioreactor simulating fed-batch cultivationusing medium limited by phosphorous, supplemented with 30 gL-1 of xylose, at pH 6.8 with different oxygen saturations. Note thatyield calculations were performed in terms of total of sugars consumed.

Cultivation XylAc P(3HB)% P(3HB)

(g P(3HB)/g CDW)

K: 20% DO

YP/S (g/g sugars) 0.01 0.000

ProdP/S (g L-1 h-1) 0.11 0.000 0

Prodmax P/S (g L-1 h-1) 0.35 0.029

L: 1% DO

YP/S (g/g sugars) - 0.149

ProdP/S (g L-1 h-1) - 0.052 30

Prodmax P/S (g L-1 h-1) - 0.060

64

glucose in the culture does not necessarily promote polymer formation by itself, or limit xylose consump-

tion. However, both shake flask and bioreactor assays indicate that glucose is preferably metabolized

compared to xylose.

Just as previous assays showed (see figure 5.15), cellular growth attained in the presence of 20% of

oxygen saturation was higher than when the oxygen supply was lower (see figures 5.17 or 5.18, for 1%

of dissolved oxygen). Xylonic acid was difficult to quantify in these assays since other HPLC equipment

was used. However, in the cultivation limited both by phosphorus and oxygen no xylonic acid production

was observed.

Previous assays showed xylonic acid production for P limited media when using 20% saturation of

DO in the medium. The amount, and therefore the yield, of xylonic acid decreased substantially when

the oxygen supply was reduced. This was to be expected since xylonic acid is produced along an

oxidative pathway.

Finally, as shown in figure 5.24 and described in table 5.8, low oxygen supply did promote polymer

production in the culture simulating the third bioreactor of the continuous setup. In fact, this assay

resulted in the highest yield attained for P(3HB) production throughout this work. This suggests that

the oxygen availability in culture medium is the target factor for enhancing P(3HB) production in xylose-

cultivations of B. sacchari .

5.2.6 B. sacchari cultivation strategies towards xylonic acid, xylitol and P(3HB)production

Table 5.9 lists the maximum yields attained in B. sacchari fed-batch cultivations within all the oper-

ation conditions studied throughout this work, allowing for an overview of what appears to be the best

cultivation conditions to promote each of the three products formation.

From the assays listed in the table below, higher cellular growth is achieved in cultivation G when

phosphorus is the limiting nutrient (ca. 47 gL-1 of residual biomass). Thus, if a cultivation is carried out

with the aim of producing a dense cell culture, P-limited medium and with 20% saturation of DO should

be used.

Xylonic acid (XylAc) seems to attain higher concentration values, and higher yield, in seeding medium

cultivations A and B, where the limiting nutrient is nitrogen. However, a yield of 0.65 g of xylonic acid

per g of xylose consumed was attained in cultivation C, medium being limited by phosphorous and with

20% of oxygen saturation.

Within this experimental work, it was not possible to conclude about the culture conditions that pro-

mote higher xylitol production. However, xylitol formation was indeed confirmed when a high xylose

concentration (above 40 gL-1) was measured in the broth.

Despite the inconclusive results regarding xylonic acid and xylitol, it was possible to conclude that

the target factor towards enhancing P(3HB) production is the oxygen availability during Burkholderia

sacchari cultivations. The results obtained suggest that if a double limitation takes place, polymer pro-

duction is induced.

65

Table 5.9: Summary of the results obtained for B. sacchari bioreactor cultivations strategies towards xylonic acid (XylAc), xylitol (Xylt) and/or P(3HB) production throughout thiswork. Residual biomass (Xr) and maximum yield obtained for each metabolite under the various operation conditions studied are shown.

Cult.Feed Limit. nutrient %DO Xr YXylAc/Xyl YXylt/Xyl YP(3HB)/Xyl %P(3HB)

Fig.Xyl Xyl+Gluc P N 1 20 (gL-1) (g/g Xylcons) (g/g Xylcons) (g/g Xylcons) (g P(3HB)/g CDW)

Ai − − + + 3.6 0.95 0.06 0.04 39 5.13

C + + + 32 0.65 − 0.00 0 5.15

D + + + 7.5 1.07 − 0.04 35 5.16

E + + + 24 0.17 0.01 0.07 26 5.17

G + + + 8.0 − − 0.12 36 5.19

K + + + 27 0.01ii − 0.00ii 0 5.23

L + + + 18 − − 0.15ii 30 5.24i Batch cultivationii g/g total sugars consumed+ states for presence− states for absence

66

6Conclusions and future prospects

67

Since the first report of a PHA-producing strain in the 1920s’, considerable progress has been ac-

complished in order to optimise the production of this polymer family, given its remarkable properties

and potential as an alternative to petroleum-based plastics. However, the high production cost of PHAs

by means of fermentative processes is the major drawback for the shift towards a more environmentally

friendly material. To overcome this issue, one of the aspects that researchers have been focusing more

recently is the use of lignocellulosic waste materials as substrate for bacterial growth, within the con-

cept of a biorefinery. In this context, special attention is being given to xylose, a major component of

lignocellulosic biomass. Moreover, two by-products of the biological conversion of xylose by Burkholde-

ria sacchari were recently reported in studies simulating lignocellulosic hydrolysates: xylitol and xylonic

acid, two value-added chemical compounds.

The work developed in this thesis aimed at finding the cultivation conditions that promote the metabolic

pathways in B. sacchari towards P(3HB), D-xylitol or D-xylonic acid production, using D-xylose as sole

carbon source. To do so, preliminary shake flask assays were performed in order to evaluate the influ-

ence of the sugar used in the inoculum cultivations, as well as the influence of yeast extract as nitrogen

source. At bioreactor scale, studies were performed to evaluate the effects of the type of nutrient used

to impose limitation and the concentration of dissolved oxygen in the medium. Confirmation of xylitol

production was also assessed in both cultivation modes.

The first shake flask assays showed that the sugar used to grow B. sacchari inocula do not greatly

influence further xylose-cultivations in terms or residual biomass and overall product yields and produc-

tivities. This means that the xylose catabolic pathways aren’t supressed by the sugar used to grow the

inoculum. With sucrose grown inocula, xylose-cultivations attained a yield of 0.09 g of polymer per g

of xylose consumed, as well as 44% of polymer content, the highest values attained in cultivations on

xylose.

The shake flask cultivations carried out to assess the effect of the yeast extract in the seeding

medium, revealed that this source of nutrients promotes cell growth. Also, higher polymer production on

xylose is attained using complete seeding medium. In cultivations carried out on xylose, the by-products

yields lie within the range 0.54-0.83 g of xylonic acid per g of xylose consumed and 0.8-0.9 g of P(3HB)

per g of xylose consumed, with a polymer content of 40%, on average.

With regard to the nutrient used to impose limitation and thus promote P(3HB) formation when xy-

lose is used as carbon source, the findings allowed to conclude that a cultivation medium limited by

nitrogen leads to higher polymer accumulation, 0.04 g of P(3HB) per g of xylose consumed (ca. 35% of

polymer content in bacterial cells), whereas phosphorus limiting conditions conduct the substrate almost

exclusively towards cell growth, reaching 32 gL-1 of residual biomass, as compared to 7.5 gL-1 under the

N-limited conditions.

Additionally to limitation of an essential nutrient supplied by the liquid phase, the findings described in

section 5.2.3 revealed that restricting the amount of oxygen available in the cultivation medium, creating

conditions for a double limitation, enhance polymer production when the medium is limited by nitrogen,

68

and also promote polymer accumulation when phosphorus is the nutrient used to impose limitation. In

these assays, the highest polymer yield was 0.12 g of P(3HB) per g of xylose consumed, for a xylose

cultivation carried out in N-limited medium with low supply of O2 (DO set-point of 1% saturation). Under

phosphorus and O2 limitation, polymer production showed a yield of 0.07 g of P(3HB) per g of xylose

consumed (corresponding to a polymer content of ca. 26%).

It was observed that the production of xylonic acid decreased by lowering the supply of oxygen to

the medium. Moreover, an overview of all the bioreactor assays allow to note that xylonic acid is very

often produced when xylose concentrations in the medium are limiting, i.e. close to zero. To confirm this

assumption, further fed-batch cultivations must be performed in order to verify if a double limitation by

both the essential nutrient and the carbon source leads to this by-product formation. In fact, the highest

yield attained for this by-product was 1.07 g of xylonic acid per g of xylose consumed in a fed-batch

cultivation carried under nitrogen limitation and concentrations of xylose in the medium close to zero.

Xylitol production was indeed confirmed for a xylose concentration in the medium of 50 gL-1, as

Raposo et al. (2017) recently reported, in both shake flask and bioreactor cultivation modes. The results

obtained for this study showed lower yields attained for P(3HB) and xylonic acid when xylitol production

occurred. This is consistent with the fact that when xylose is being converted into xylitol less substrate

is used by the cell to produce polymer and xylonic acid.

Even though, according to Himabindu and Gummadi (2015), at high xylose concentrations, growth

and product formation are inhibited due to osmotic stress, in B. sacchari metabolism a high sugar con-

centration seems to channel xylose conversion towards xylitol formation. The reasons for this are yet

unknown. However, observing the metabolic network proposed (see figure 2.5), a possible explanation

could be that the xylitol formation pathway is the quickest route for the cell to convert xylose and thus

counteract the osmotic pressure of the extracellular medium.

Despite the confirmation of xylitol production, it was not possible to achieve the cultivation conditions

that promote this by-product formation and withdraw relevant conclusions. Further studies must be

performed in order to enhance xylitol production and yield. Moreover, the oxygen supply should also be

assessed in these cultivations. In their review, Himabindu and Gummadi (2015) mention that oxygen

limitation may increase xylitol production since xylitol dehydrogenase (XOHDH) is inhibited due to NADH

regeneration impairment which results in a high NADH/NAD+ ratio. Therefore, a double limitation may

enhance both xylitol and polymer production.

Another approach proposed by Himabindu and Gummadi (2015) to enhance xylitol formation is the

study of the presence of glucose in the medium as co-substrate: by supplementing this hexose in the

feed, it will be uptaken for cell growth, whereas only a small amount of xylose will be used for cell growth

and maintenance, and therefore the major fraction of this pentose will be channeled into xylitol formation,

thus enhancing the yield attained.

Finally, the assay described in section 5.2.5, simulating the third bioreactor of the continuous culti-

vation system run by Henriques (2015), showed that limiting the oxygen supply in the third bioreactor

69

would implicate a double limitation in the broth, thus promoting polymer formation by B. sacchari on

xylose. These findings allow to confirm that the target factor for P(3HB) production by B. sacchari on

xylose is indeed the amount of oxygen available during the time course of cultivations. In this assay,

a yield of 0.15 g of P(3HB) per g of total sugars consumed was attained, corresponding to a polymer

content of 30%.

In short, with the work performed throughout this thesis it is possible to conclude that, in B. sacchari

cultivations, the oxygen supply is one possible target factor to enhance P(3HB) production, and also that

the xylose concentration present in the broth is very likely the target factor for both xylitol and xylonic

acid production improvement.

Concerning future work, it is of upmost importance the necessity of statistical significance regarding

all the assays performed. The conclusions discussed throughout this document are mainly assumptions

regarding the experimental results obtained and literature review. Therefore, further studies are required

in order to confirm these findings and report them.

Additionally, the knowledge of the metabolic network operation of B. sacchari by means of metabolic

flux analysis, will enable a rational streamlining of the catabolic pathways that lead to an efficient carbon

source utilization, and thus improve polymer and other metabolites production and accumulation. For

this assessment total protein quantification it is required, therefore cell disruption studies must also be

performed.

Furthermore, other aspects regarding the experimental equipments may improve results in the fu-

ture. For instance, online xylose, phosphate and ammonia sensors could be used to achieve the ”opti-

mal” cultivation conditions in terms of substrate and nutrients supply (or limitation). Also, volume level

sensors and an automatic sampler would allow for more accurate calculations and therefore yield and

productivities values achieved.

As final remarks, although not all the objectives of the study proposed in the beginning of this thesis’

were fully met, the outcome of the work developed provided two possible guidelines for P(3HB), xylitol

and xylonic acid production improvement. Further research exploring how these target factors can be

used to achieve high productivities will allow a more efficient pentose consumption. The latter is a topic

of upmost importance in the upgrade of lignocellulosic biomass and may have a large impact on the

progress of LCF-biorefineries, which might be a contribuition to a more sustainable society.

70

Bibliography

Alexandrino, P. M. R., Mendonca, T. T., Guaman Bautista, L. P., Cherix, J., Lozano-Sakalauskas, G. C.,

Fujita, A., Ramos Filho, E., Long, P., Padilla, G., Taciro, M. K., Gomez, J. G. C., and Silva, L. F. (2015).

Draft Genome Sequence of the Polyhydroxyalkanoate-Producing Bacterium Burkholderia sacchari

LMG 19450 Isolated from Brazilian Sugarcane Plantation Soil. Genome Announc., 3(3):e00313–15.

Bertrand, J.-L., Ramsay, B., Ramsay, J., and Chavarie, C. (1990). Biosynthesis of Poly-β-

Hydroxyalkanoates from Pentoses by Pseudomonas pseudoflava. Appl. Environ. Microbiol.,

56(10):3133–3138.

Bramer, C. O., Vandamme, P., da Silva, L. F., Gomez, J., and Steinbuchel, A. (2001). Burkholderia

sacchari sp. nov., a polyhydroxyalkanoate-accumulating bacterium isolated from soil of a sugar-cane

plantation in Brazil. Int. J. Syst. Evol. Microbiol., 51(5):1709–1713.

Buchert, J., Puls, J., and Poutanen, K. (1988). Comparison of Pseudomonas fragi and Gluconobacter

oxydans for production of xylonic acid from hemicellulose hydrolyzates. Appl. Microbiol. Biotechnol.,

28(4-5):367–372.

Buchert, J. and Viikari, L. (1988). The role of xylonolactone in xylonic acid production by Pseudomonas

fragi . Appl. Microbiol. Biotechnol., 27(4):333–336.

CalRecycle (2012). Department of Resources Recycling and Recovery.

Cavalheiro, J. M., Raposo, R. S., de Almeida, M. C. M., Teresa Cesario, M., Sevrin, C., Grandfils, C., and

da Fonseca, M. (2012). Effect of cultivation parameters on the production of poly(3-hydroxybutyrate-

co-4-hydroxybutyrate) and poly(3-hydroxybutyrate-4-hydroxybutyrate-3-hydroxyvalerate) by Cupri-

avidus necator using waste glycerol. Bioresour. Technol., 111:391–397.

Cesario, M. T., Raposo, R. S., de Almeida, M. C. M., van Keulen, F., Ferreira, B. S., and da Fonseca,

M. M. R. (2014). Enhanced bioproduction of poly-3-hydroxybutyrate from wheat straw lignocellulosic

hydrolysates. N. Biotechnol., 31(1):104–113.

Cesario, M. T. F. and de Almeida, M. C. M. D. (2015). Lignocellulosic Hydrolysates for the Production

of Polyhydroxyalkanoates. In Kamm, B., editor, Microorg. Biorefineries, volume 26, chapter 4, pages

79–104. Springer-Verlag Berlin Heidelberg, Berlin, first edition.

Chanprateep, S. (2010). Current trends in biodegradable polyhydroxyalkanoates. J. Biosci. Bioeng.,

110(6):621–632.

71

Chen, X. (2010). Microbial and Bioconversion Production of D-xylitol and Its Detection and Application.

Int. J. Biol. Sci., 6(7):834–844.

Cherix, J., Guaman, L., Gomez, J., Taciro, M., and Silva, L. (2014). Analysis of xylose catabolism in

Burkholderia sacchari for the production of polyhydroxyalkanoates. 14th Int. Symp. Biopolym.

da Fonseca, M. M. R. and Teixeira, J. A. (2007). Reactores Biologicos - Fundamentos e Aplicacoes.

LIDEL - Edicoes Tecnicas, Lda., Lisbon, first edition.

de Jong, E. and Jungmeier, G. (2015). Biorefinery Concepts in Comparison to Petrochemical Refineries.

In Ind. Biorefineries White Biotechnol., pages 3–33. Elsevier.

De Lederkremer, R. M. and Marino, C. (2004). Acids and Other Products of Oxidation of Sugars. Adv.

Carbohydr. Chem. Biochem., 58(36):199–306.

Dias, J. M. L., Lemos, P. C., Serafim, L. S., Oliveira, C., Eiroa, M., Albuquerque, M. G. E., Ramos, A. M.,

Oliveira, R., and Reis, M. A. M. (2006). Recent Advances in Polyhydroxyalkanoate Production by

Mixed Aerobic Cultures: From the Substrate to the Final Product. Macromol. Biosci., 6(11):885–906.

Doan, T. V. and Nguyen, B. T. (2012). Polyhydroxyalkanoates production by a bacterium isolated from

mangrove soil samples collected from Quang Ninh province. J. Vietnamese Environ., 3(2):76–79.

FitzPatrick, M., Champagne, P., Cunningham, M. F., and Whitney, R. A. (2010). A biorefinery process-

ing perspective: Treatment of lignocellulosic materials for the production of value-added products.

Bioresour. Technol., 101(23):8915–8922.

Frost, J. (2008). Creating a pathway for the biosynthesis of 1,2,4-butanetriol. Inter-Agency Conf. Metab.

Eng.

Gomez, J., da Silva, L., and Mendonca, T. (2014). Characterization of bacterial platforms based on PHA

production analysis. 14th Int. Symp. Biopolym.

Gottschalk, G. (1986). Bacterial Metabolism. Springer Series in Microbiology. Springer New York, New

York, NY.

Gross, R. A. (2002). Biodegradable Polymers for the Environment. Science, 297(5582):803–807.

Halami, P. M. (2008). Production of polyhydroxyalkanoate from starch by the native isolate Bacillus

cereus CFR06. World J. Microbiol. Biotechnol., 24(6):805–812.

Henriques, C. P. (2015). Towards Polyhydroxyalkanoates Production in Three-stage Continuous Biore-

actors using Glucose/Xylose Mixtures Simulating Lignocellulosic Hydrolysates. Master thesis, Instituto

Superior Tecnico.

Hernandez-Perez, A., Costa, I., Silva, D., Dussan, K., Villela, T., Canettieri, E., Carvalho, J., Soares

Neto, T., and Felipe, M. (2016). Biochemical conversion of sugarcane straw hemicellulosic hydrolyzate

supplemented with co-substrates for xylitol production. Bioresour. Technol., 200(1):1085–1088.

72

Himabindu, K. and Gummadi, S. (2015). Effect of kLa and Fed-batch Strategies for Enhanced Production

of Xylitol by Debaryomyces nepalensis NCYC 3413. Br. Biotechnol. J., 5(1):24–36.

Ienczak, J. L., Schmidell, W., and de Aragao, G. M. F. (2013). High-cell-density culture strategies for

polyhydroxyalkanoate production: a review. J. Ind. Microbiol. Biotechnol., 40(3-4):275–286.

Jeffries, T. W. (1983). Utilization of xylose by bacteria, yeasts, and fungi. In Pentoses and Lignin,

volume 27 of Advances in Biochemical Engineering/Biotechnology, pages 1–32. Springer Berlin Hei-

delberg, Berlin.

Kamm, B. and Kamm, M. (2004). Principles of biorefineries. Appl. Microbiol. Biotechnol., 64(2):137–145.

Kamm, B. and Kamm, M. (2007). Biorefineries – Multi Product Processes. In Ulber, R. and Sell, D.,

editors, White Biotechnol., chapter 5, pages 175–204. Springer Berlin Heidelberg, Berlin, Heidelberg.

Kanehisa, M., Sato, Y., Kawashima, M., Furumichi, M., and Tanabe, M. (2016). KEGG as a reference

resource for gene and protein annotation. Nucleic Acids Res., 44(D1):D457–D462.

Kasprow, R., Lange, A., and Kirwan, D. (1998). Correlation of Fermentation Yield with Yeast Extract

Composition as Characterized by Near-Infrared Spectroscopy. Biotechnol. Prog., 14(2):318–325.

Kim, B. S., Lee, S. C., Lee, S. Y., Chang, H. N., Chang, Y. K., and Woo, S. I. (1994). Production of

poly(3-hydroxybutyric acid) by fed-batch culture of Alcaligenes eutrophus with glucose concentration

control. Biotechnol. Bioeng., 43(9):892–898.

Kwon, S.-G., Park, S.-W., and Oh, D.-K. (2006). Increase of xylitol productivity by cell-recycle fermenta-

tion of Candida tropicalis using submerged membrane bioreactor. J. Biosci. Bioeng., 101(1):13–18.

Lam, V. M., Daruwalla, K. R., Henderson, P. J., and Jones-Mortimer, M. C. (1980). Proton-linked D-

xylose transport in Escherichia coli . Jounal Bacteriol., 143(1):396–402.

Lee, S. Y. (1996). Bacterial polyhydroxyalkanoates. Biotechnol. Bioeng., 49(1):1–14.

Lee, S. Y. (1998). Poly (3-hydroxybutyrate) production from xylose by recombinant Escherichia coli.

Bioprocess Eng., 18:397–399.

Liu, H., Valdehuesa, K. N. G., Nisola, G. M., Ramos, K. R. M., and Chung, W.-J. (2012). High yield

production of d-xylonic acid from d-xylose using engineered Escherichia coli . Bioresour. Technol.,

115:244–248.

Lopes, M. S. G., Gosset, G., Rocha, R. C. S., Gomez, J. G. C., and Ferreira da Silva, L. (2011). PHB

Biosynthesis in Catabolite Repression Mutant of Burkholderia sacchari . Curr. Microbiol., 63(4):319–

326.

Lopes, M. S. G., Rocha, R. C. S., Zanotto, S. P., Gomez, J. G. C., and da Silva, L. F. (2009). Screening of

bacteria to produce polyhydroxyalkanoates from xylose. World J. Microbiol. Biotechnol., 25(10):1751–

1756.

73

Lopez, F., Delgado, O. D., Martınez, M. A., Spencer, J. F., and Figueroa, L. I. (2004). Characterization

of a new xylitol-producer Candida tropicalis strain. Antonie Van Leeuwenhoek, 85(4):281–286.

Minghua, Z., Xiumin, F., Rovetta, A., Qichang, H., Vicentini, F., Bingkai, L., Giusti, A., and Yi, L. (2009).

Municipal solid waste management in Pudong New Area, China. Waste Manag., 29(3):1227–1233.

Moat, A. G., Foster, J. W., and Spector, M. P. (2002). Microbial Physiology. John Wiley & Sons, Inc.,

Hoboken, NJ, USA, 4 edition.

Niu, W., Molefe, M. N., and Frost, J. W. (2003). Microbial Synthesis of the Energetic Material Precursor

1,2,4-Butanetriol. J. Am. Chem. Soc., 125(43):12998–12999.

Nygard, Y., Toivari, M. H., Penttila, M., Ruohonen, L., and Wiebe, M. G. (2011). Bioconversion of D-

xylose to D-xylonate with Kluyveromyces lactis. Metab. Eng., 13(4):383–391.

Ogata, H., Goto, S., Sato, K., Fujibuchi, W., Bono, H., and Kanehisa, M. (1999). KEGG: Kyoto Encyclo-

pedia of Genes and Genomes. Nucleic Acids Res., 27(1):29–34.

Pezzotti, F. and Therisod, M. (2006). Enzymatic synthesis of aldonic acids. Carbohydr. Res.,

341(13):2290–2292.

PNNL and NREL (2004). Top Value Added Chemicals from Biomass: Volume 1 - Results of Screening for

Potential Candidates from Sugars and Synthesis Gas. Technical report, U.S. Department of Energy.

Radek, A., Krumbach, K., Gatgens, J., Wendisch, V. F., Wiechert, W., Bott, M., Noack, S., and Marien-

hagen, J. (2014). Engineering of Corynebacterium glutamicum for minimized carbon loss during

utilization of d-xylose containing substrates. J. Biotechnol., 192, Part:156–160.

Rafiqul, I. S. M. and Mimi Sakinah, A. M. (2012). A perspective: Bioproduction of xylitol by enzyme

technology and future prospects. Int. Food Res. J., 19(2):405–408.

Ragauskas, A. J. (2006). The Path Forward for Biofuels and Biomaterials. Science, 311(5760):484–489.

Raposo, R. S., de Almeida, M. C. M., de Oliveira, M. d. C. M., da Fonseca, M. M., and Cesario, M. T.

(2017). A Burkholderia sacchari cell factory: production of poly-3-hydroxybutyrate, xylitol and xylonic

acid from xylose-rich sugar mixtures. N. Biotechnol., 34:12–22.

Ravella, S. R., Gallagher, J., Fish, S., and Prakasham, R. S. (2012). Overview on Commercial Produc-

tion of Xylitol, Economic Analysis and Market Trends. In D-Xylitol, pages 291–306. Springer Berlin

Heidelberg, Berlin, Heidelberg.

Roberto, I. C., Felipe, M. G., de Mancilha, I. M., Vitolo, M., Sato, S., and da Silva, S. S. (1995). Xyli-

tol production by Candida guillermondii as an approach for the utilization of agroindustrial residues.

Bioresour. Technol., 51(2-3):255–257.

Ryu, H. W., Hahn, S. K., Chang, Y. K., and Chang, H. N. (1997). Production of poly(3-hydroxybutyrate)

by high cell density fed-batch culture of Alcaligenes eutrophus with phosphate limitation. Biotechnol.

Bioeng., 55(1):28–32.

74

SIADEB (2010). SIADEB Brochure.

Silva, L. F., Taciro, M. K., Michelin Ramos, M. E., Carter, J. M., Pradella, J. G. C., and Gomez, J. G. C.

(2004). Poly-3-hydroxybutyrate (P3HB) production by bacteria from xylose, glucose and sugarcane

bagasse hydrolysate. J. Ind. Microbiol. Biotechnol., 31(6):245–254.

Song, Q., Li, J., and Zeng, X. (2015). Minimizing the increasing solid waste through zero waste strategy.

J. Clean. Prod., 104:199–210.

Stephen Dahms, A. (1974). 3-Deoxy-D-pentulosonic acid aldolase and its role in a new pathway of

D-xylose degradation. Biochem. Biophys. Res. Commun., 60(4):1433–1439.

Sudesh, K., Abe, H., and Doi, Y. (2000). Synthesis, structure and properties of polyhydroxyalkanoates:

biological polyesters. Prog. Polym. Sci., 25(10):1503–1555.

Suriyamongkol, P., Weselake, R., Narine, S., Moloney, M., and Shah, S. (2007). Biotechnological ap-

proaches for the production of polyhydroxyalkanoates in microorganisms and plants — A review.

Biotechnol. Adv., 25(2):148–175.

Toivari, M., Nygard, Y., Kumpula, E.-P., Vehkomaki, M.-L., Bencina, M., Valkonen, M., Maaheimo, H.,

Andberg, M., Koivula, A., Ruohonen, L., Penttila, M., and Wiebe, M. G. (2012). Metabolic engineering

of Saccharomyces cerevisiae for bioconversion of D-xylose to D-xylonate. Metab. Eng., 14(4):427–

436.

Toivari, M., Vehkomaki, M.-L., Nygard, Y., Penttila, M., Ruohonen, L., and Wiebe, M. G. (2013). Low pH

D-xylonate production with Pichia kudriavzevii . Bioresour. Technol., 133:555–562.

United Nations Department of Economic and Social Affairs Population Division (2015). World Population

Prospects: The 2015 Revision, Key Findings and Advance Tables. Technical report, United Nations,

New York.

van Ree, R. and van Zeeland, A. (2014). IEA Bioenergy: Task42 Biorefining. pages 1–66.

Venkateswar Rao, L., Goli, J. K., Gentela, J., and Koti, S. (2016). Bioconversion of lignocellulosic

biomass to xylitol: An overview. Bioresour. Technol., 213:299–310.

Verlinden, R., Hill, D., Kenward, M., Williams, C., and Radecka, I. (2007). Bacterial synthesis of

biodegradable polyhydroxyalkanoates. J. Appl. Microbiol., 102(6):1437–1449.

Wang, C., Wei, D., Zhang, Z., Wang, D., Shi, J., Kim, C. H., Jiang, B., Han, Z., and Hao, J. (2016).

Production of xylonic acid by Klebsiella pneumoniae. Appl. Microbiol. Biotechnol., 100(23):10055–

10063.

Weimberg, R. (1961). Pentose oxidation by Pseudomonas fragi. J. Biol. Chem., 236(3):629–635.

Winkelhausen, E. and Kuzmanova, S. (1998). Microbial conversion of d-xylose to xylitol. J. Ferment.

Bioeng., 86(1):1–14.

75

Yoshitake, J., Ishizaki, H., Shimamura, M., and Imai, T. (1973). Xylitol Production by an Enterobacter

Species. Agric. Biol. Chem., 37(10):2261–2267.

Yoshitake, J., Ohiwa, H., Shimamura, M., and Imai, T. (1971). Production of Polyalcohol by a Corynebac-

terium sp. Agric. Biol. Chem., 35(6):905–911.

Young, F. K., Kastner, J. R., and May, S. W. (1994). Microbial production of poly-β-hydroxybutyric acid

from d-xylose and lactose by Psudomonas cepacia. Appl. Environ. Microbiol., 60(11):4195–4198.

Zaman, A. U. (2015). A comprehensive review of the development of zero waste management: lessons

learned and guidelines. J. Clean. Prod., 91:12–25.

Zero Waste International Alliance (2009). Zero Waste Definition.

Zhang, J., Geng, A., Yao, C., Lu, Y., and Li, Q. (2012). Xylitol production from D-xylose and horticultural

waste hemicellulosic hydrolysate by a new isolate of Candida athensensis SB18. Bioresour. Technol.,

105:134–141.

Zhou, X., Wang, X., Cao, R., Tao, Y., Xu, Y., and Yu, S. (2015). Characteristics and Kinetics of the Aldonic

Acids Production using Whole-cell catalysis of Gluconobacter oxydans. BioResources, 10(3):4277–

4286.

76

ASubstrates, by-products of xylose

metabolism, and ammoniumquantification

A-1

A.1 Carbon sources, xylonic acid, xylitol and phosphate determi-nations

Equations A.1 to A.6 describe the calibration curves obtained, as well as the corresponding correla-

tion factor, for glucose, xylose, xylonic acid, xylitol and phosphate determinations, for working ranges of

0.5 to 60 gL-1 for glucose and xylose, 0.5 to 60 gL-1 for xylonic acid, 0.1 to 5 gL-1 for xylitol, and 0.5 to

60 gL-1 for phosphate, respectively.

[Glucose] (gL−1) = 7.89 × 10−6 ×Apeak − 2.10 × 10−1 (r2 = 0.9990) (A.1)

[Xylose] (gL−1) = 7.92 × 10−6 ×Apeak − 5.22 × 10−2 (r2 = 0.9996) (A.2)

[Xylonicacid]UV (gL−1) = 5.09 × 10−6 ×Apeak − 3.13 × 10−1 (r2 = 0.9988) (A.3)

[Xylonicacid]IR (gL−1) = 1.20 × 10−5 ×Apeak − 1.32 × 100 (r2 = 0.9941) (A.4)

[Xylitol] (gL−1) = 8.00 × 10−6 ×Apeak − 2.13 × 10−1 (r2 = 0.9993) (A.5)

[Phosphate] (gL−1) = 1.00 × 10−5 ×Apeak − 9.97 × 10−2 (r2 = 0.9995) (A.6)

Equations A.7 to A.12 were obtained as the calibration curves for the second HPLC used to deter-

mine carbon sources, xylonic acid, xylitol and phosphate.

[Glucose] (gL−1) = 8.46 × 10−6 ×Apeak + 1.21 × 10−1 (r2 = 0.9957) (A.7)

[Xylose] (gL−1) = 8.03 × 10−6 ×Apeak + 4.06 × 10−2 (r2 = 0.9912) (A.8)

[Xylonicacid]UV (gL−1) = 6.13 × 10−6 ×Apeak − 6.3 × 100 (r2 = 0.9510) (A.9)

[Xylonicacid]RI (gL−1) = 1.42 × 10−5 ×Apeak − 4.9 × 100 (r2 = 0.9638) (A.10)

[Xylitol] (gL−1) = 6.00 × 10−6 ×Apeak + 4.60 × 10−1 (r2 = 0.9920) (A.11)

[Phosphate] (gL−1) = 2.00 × 10−5 ×Apeak − 3.93 × 10−1 (r2 = 0.9856) (A.12)

Since xylitol and phosphate had the similar retention times with this HPLC apparatus, a new calibra-

tion curve for xylitol determination was performed to overcome this issue, which is described in equation

A.11.

A-2

Figure A.1: HPLC chromatogram for RI detector.

Figure A.2: HPLC chromatogram for UV-visible detector.

A-3

A.2 P(3HB) determination

The calibration curve for P(3HB) determination was determined for a working range of 0 to 10 gL-1,

and is described in equation A.13.

[P (3HB)] (gL−1) = 8.84 × 100 ×(AP (3HB) peak/AIS peak

)+ 2.34 × 10−1 (r2 = 0.998) (A.13)

A.3 Ammonium determination

In order assess the availability of nitrogen in the broth and thus determine the ammonium concen-

tration, conductivity measurements were performed in supernatant of cultivation samples carried out in

N-limited medium, as described in section 4.6.5. Therefore, equation A.14 to A.16 describe the calibra-

tion curves used to determine the ammonium concentration for cultivations D, G and H, correspondig to

figures 5.16, 5.19 and 5.20, respectively. Note that the concentration computed through the calibration

curves is in M. Considering the molar mass of the ammonium, the concentration is then given in gL-1.

Cond (mV ) = 44.07 × log [NH4+] + 149.41 (r2 = 0.9706) (A.14)

Cond (mV ) = 33.73 × log [NH4+] + 107.94 (r2 = 0.9940) (A.15)

Cond (mV ) = 44.95 × log [NH4+] + 170.55 (r2 = 0.9933) (A.16)

A-4


Recommended