genome-scale metabolic modelling of arabidopsis and tomato · genome-scale metabolic modelling of...

156
Genome-scale metabolic modelling of Arabidopsis and tomato Citation for published version (APA): Yuan, H. (2016). Genome-scale metabolic modelling of Arabidopsis and tomato. Eindhoven: Technische Universiteit Eindhoven. Document status and date: Published: 14/12/2016 Document Version: Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers) Please check the document version of this publication: • A submitted manuscript is the version of the article upon submission and before peer-review. There can be important differences between the submitted version and the official published version of record. People interested in the research are advised to contact the author for the final version of the publication, or visit the DOI to the publisher's website. • The final author version and the galley proof are versions of the publication after peer review. • The final published version features the final layout of the paper including the volume, issue and page numbers. Link to publication General rights Copyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright owners and it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights. • Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal. If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, please follow below link for the End User Agreement: www.tue.nl/taverne Take down policy If you believe that this document breaches copyright please contact us at: [email protected] providing details and we will investigate your claim. Download date: 05. Mar. 2020

Upload: others

Post on 03-Mar-2020

6 views

Category:

Documents


0 download

TRANSCRIPT

Genome-scale metabolic modelling of Arabidopsis andtomatoCitation for published version (APA):Yuan, H. (2016). Genome-scale metabolic modelling of Arabidopsis and tomato. Eindhoven: TechnischeUniversiteit Eindhoven.

Document status and date:Published: 14/12/2016

Document Version:Publisher’s PDF, also known as Version of Record (includes final page, issue and volume numbers)

Please check the document version of this publication:

• A submitted manuscript is the version of the article upon submission and before peer-review. There can beimportant differences between the submitted version and the official published version of record. Peopleinterested in the research are advised to contact the author for the final version of the publication, or visit theDOI to the publisher's website.• The final author version and the galley proof are versions of the publication after peer review.• The final published version features the final layout of the paper including the volume, issue and pagenumbers.Link to publication

General rightsCopyright and moral rights for the publications made accessible in the public portal are retained by the authors and/or other copyright ownersand it is a condition of accessing publications that users recognise and abide by the legal requirements associated with these rights.

• Users may download and print one copy of any publication from the public portal for the purpose of private study or research. • You may not further distribute the material or use it for any profit-making activity or commercial gain • You may freely distribute the URL identifying the publication in the public portal.

If the publication is distributed under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license above, pleasefollow below link for the End User Agreement:

www.tue.nl/taverne

Take down policyIf you believe that this document breaches copyright please contact us at:

[email protected]

providing details and we will investigate your claim.

Download date: 05. Mar. 2020

Genome-Scale Metabolic

Modelling of Arabidopsis and

Tomato

This work was supported by the Brainbridge project (collaboration between Eind-

hoven University of Technology, Zhejiang University and Philips Asia-Shanghai)

and the China Scholarship Council (CSC).

A catalogue is available from the Eindhoven University of Technology Library

ISBN: 978-90-386-4191-1

Cover Design: Xinwei Wang and Koen Pieterse

Cover image: A picture of the green leaf from Reshma Chowdhury with improved

design.

Printed by: Ipskamp Drukkers, Enschede, the Netherlands

Copyright © 2016 by Huili Yuan

Genome-Scale Metabolic

Modelling of Arabidopsis and

Tomato

PROEFSCHRIFT

ter verkrijging van de graad van doctor

aan de Technische Universiteit Eindhoven,

op gezag van de rector magnificus prof.dr.ir. F.P.T. Baaijens,

voor een commissie aangewezen door het College voor Promoties,

in het openbaar te verdedigen

op woensdag 14 december 2016 om 14:00 uur

door

Huili Yuan

geboren te Henan, China

Dit proefschrift is goedgekeurd door de promotoren en de samenstelling van de pro-

motiecommissie is als volgt:

Voorzitter: prof.dr. C.W.J. Oomens

promotor: prof.dr. P.A.J. Hilbers

copromotor: prof.dr. N.A.W. van Riel

leden: prof.dr. B. Teusink (Universiteit van Amsterdam)

prof.dr. R.G. Ratcliffe (University of Oxford)

prof.dr. J.W.M. Bergmans

prof.dr. G.F. Zhou (TU/e; South China Normal University)

adviseur: dr. M.C.Y. Cheung (Yale-NUS College)

Het onderzoek of ontwerp dat in dit proefschrift wordt beschreven is uitgevoerd in

overeenstemming met de TU/e Gedragscode Wetenschapsbeoefening.

I

Contents

Chapter 1 Introduction .............................................................................. 1

1.1 Plant metabolism and its investigation .................................................... 2

1.2 Constraint-based modelling ..................................................................... 4

1.2.1 Flux balance analysis.................................................................. 5

1.2.2 Flux variability analysis ............................................................. 6

1.2.3 Flux-sum analysis ....................................................................... 6

1.3 Genome-scale metabolic models ............................................................. 6

1.3.1 Reconstruction of GSMs ............................................................ 8

1.3.2 Software for genome-scale metabolic modelling ....................... 9

1.3.3 Applications of GSMs .............................................................. 10

1.4 Organization of this thesis ..................................................................... 11

Chapter 2 Flux balance analysis of plant metabolism: the effect of

biomass composition and model structure on model predictions .............. 13

2.1 Introduction ........................................................................................... 15

2.2 Methods ................................................................................................. 17

2.2.1 Stoichiometric models .............................................................. 17

2.2.2 Biomass equations .................................................................... 18

2.2.3 Model simulations .................................................................... 18

2.3 Results… ............................................................................................... 21

2.3.1 Comparisons between plant flux-balanced models .................. 21

Contents

II

2.3.2 The impact of biomass composition and model structure on

central metabolic fluxes............................................................ 26

2.4 Discussion ............................................................................................. 31

2.5 Conclusions ........................................................................................... 32

2.6 Appendix ............................................................................................... 33

Chapter 3 Constructing a genome-scale metabolic model for

tomato…………….. ....................................................................................... 39

3.1 Introduction ........................................................................................... 41

3.2 Methods ................................................................................................. 42

3.2.1 Software ................................................................................... 42

3.2.2 Metabolic reconstruction .......................................................... 42

3.2.3 Model analysis .......................................................................... 43

3.3 Results and Discussion .......................................................................... 45

3.3.1 General model properties ......................................................... 45

3.3.2 Comparisons with existing plant models .................................. 46

3.3.3 Model validation ...................................................................... 47

3.4 Conclusions ........................................................................................... 52

3.5 Appendix ............................................................................................... 52

Chapter 4 In silico study of photorespiratory metabolism under

drought conditions…………….. ................................................................... 57

4.1 Introduction ........................................................................................... 59

4.2 Methods ................................................................................................. 60

4.2.1 Model analysis .......................................................................... 60

4.2.2 Deletion simulation .................................................................. 60

4.3 Results and Discussion .......................................................................... 60

Contents

III

4.3.1 Differences in metabolically active reactions between normal

and drought conditions ............................................................. 61

4.3.2 Effect of drought stress on the core metabolic pathways ......... 66

4.3.3 Evaluation of reaction deletions ............................................... 68

4.4 Conclusions ........................................................................................... 72

4.5 Appendix ............................................................................................... 72

Chapter 5 Modelling the effects of light intensity and quality on light-

driven metabolism of tomato ........................................................................ 77

5.1 Introduction ........................................................................................... 79

5.2 Methods ................................................................................................. 81

5.2.1 Reconstruction of a tomato light-specific model ...................... 81

5.2.2 Constraint-based flux analysis .................................................. 82

5.2.3 Flux-sum ................................................................................... 83

5.3 Results……….. ..................................................................................... 83

5.3.1 Representation of photosynthetic light reactions in existing plant

GSMs ........................................................................................ 83

5.3.2 Two published light-specific models........................................ 85

5.3.3 Model properties ...................................................................... 87

5.3.4 Growth simulation of tomato cells under different light

conditions ................................................................................. 88

5.3.5 Exploring energy balancing in the light ................................... 90

5.3.6 Flux-sum differences in cofactors ATP, NADH and

NADPH……………………………………………………….93

5.4 Discussion ............................................................................................. 94

5.4.1 Influence of varying light on the cellular growth ..................... 94

5.4.2 Energy balancing under varying light conditions ..................... 95

5.4.3 Potential limitations .................................................................. 96

Contents

IV

5.5 Conclusions ........................................................................................... 98

5.6 Appendix ............................................................................................... 99

Chapter 6 General discussion ................................................................ 103

6.1 Summary of contributions ................................................................... 104

6.2 Limitations of this study ...................................................................... 104

6.3 Future perspective ............................................................................... 106

References………… ........................................................................................... 109

Summary………………… ................................................................................. 135

Publications……….. ........................................................................................... 137

Acknowledgements ............................................................................................. 138

Curriculum Vitae…………. .............................................................................. 141

V

Abbreviations

2-OG α-ketoglutarate

2-PG 2-phosphoglycolate

2-PGA 2-phosphoglycerate

3-PGA 3-phosphoglycerate

6PG 6-phosphogluconate

6PGD 6-Phosphogluconate dehydrogenase

6PGL 6-Phosphogluconolactonase

6PGL 6-phosphogluconolactone

ABA abscisic acid

AcCoA acetyl-CoA

ACET acetate

ACETALD acetaldehyde

ADPG ADP-glucose

AEF alternate electron flow

AGT alanine-glyoxylate transaminase

Ala alanine

ALDH aldehyde dehydrogenase

AOX alternative oxidase

Arg arginine

Asn asparagine

Asp aspartate

ATPase ATP synthase

BOF biomass objective function

BRENDA Braunschweig Enzyme Database

CAT catalase

CBM constraint-based modelling

CEF cyclic electron flow

CIT citrate synthase

Cit citrate

CMP cytidine monophosphate

COBRA COnstraints Based Reconstruction And analysis

COX cytochrome C oxidase

Abbreviations

VI

Cys cysteine

Cytob6f cytochrome b6f complex

Cytox cytochrome oxidized

Cytred cytochrome reduced

DHAP dihydroxyacetone phosphate

DMAPP dimethylallyl pyrophosphate

DPG glycerate-1,3-bisphosphate

DVPCHLDOR divinyl chlorophyllide-a:NADP+ oxidoreductase

DW dry weight

e- electrons

E4P erythrose-4-phosphate

ETC electron transport chain

F6P fructose-6-phosphate

FBA flux balance analysis

FBP fructose-1,6-biphosphate

Fd-GOGAT ferredoxin-dependent glutamate synthase

Fdox ferredoxin oxidized

Fdred ferredoxin reduced

FNR ferredoxin NADP+ reductase

FQR ferredoxin plastoquinone reductase

FSA flux sum analysis

Fum fumarate

FVA flux variability ananlysis

FW fresh weight

G1P glucose-1-phosphate

G6P glucose-6-phosphate

G6PD glucose 6-phosphate dehydrogenase

GAL glutamate-ammonia ligase

GAM growth associated maintenance

GAP glyceraldehyde-3-phosphate

GAPDH glyceraldehyde-3-phosphate dehydrogenase

GDC glycine decarboxylase

GGT glutamate:glyoxylate aminotransferase

GLBE 1,4-alpha-glucan branching enzyme

GLC glucose

Gln glutamine

GLPK Gnu Linear Programming Kit

Abbreviations

VII

Glt glycolate

Glu glutamate

Gly glycine

GLYK glycerate kinase

GOGAT glutamate synthase

GOX glycolate oxidase

Gox glyoxylate

GPR gene-protein-reaction

GS glutamine synthetase

GSM genome-scale metabolic model

GTR glycine transaminase

H2O2 hydrogen peroxide

HPR hydroxypyruvate reductase

IDH isocitrate dehydrogenase

Ile isoleucine

IPP isopentenyl pyrophosphate

IsoCit threo-isocitrate

KEGG Kyoto Encyclopedia of Genes and Genomes

LED light emitting diode

LEF linear electron flow

Leu leucine

LP Linear Programming

LSP light saturation point

Lys lysine

Mal malate

MDH malate dehydrogenase

Mehler Mehler’s reaction

MEP 2C-methyl-D-erythritol-4-phosphate

Met methionine

M-THF methylene-tetrahydrofolate

MVA mevalonate

NDH NADPH dehydrogenase

NGAM non-growth associated maintenance

NiR nitrite reductase

NR nitrate reductase

OAA oxalacetic acid

OGDH 2-oxoglutarate dehydrogenase

OH-PYR hydroxypyruvate

Abbreviations

VIII

PC plastocyanin

PCHLDOR chlorophyllide-a:NADP+ oxidoreductase

PDH pyruvate dehydrogenase

PEP phosphoenol pyruvate

PEPC phosphoenolpyruvate carboxylase

PETs photosynthetic electron transport flows

PGK phosphoglycerate kinase

Phe phenylalanine

PHOTOR photorespiration

Pi inorganic phosphate

PLGP phosphoglycolate phosphatase

PPDK pyruvate orthophosphate dikinase

PPi diphosphate

PQ plastoquinone

Pro proline

PSI photosystem I

PSII photosystem II

PTOX plastoquinol oxidase

Pyr pyruvate

Q ubiquinol

QH2 ubiquinone

R5P ribose-5-phosphate

RPE ribulose-5-phosphate epimerase

RPI ribulose-5-phosphate isomerase

RPK ribulose-5-phosphate kinase

Ru5P ribulose-5-phosphate

RubisCO ribulose-1,5-bisphosphate carboxylase/oxygenase

RuBP ribulose-1,5-bisphosphate

S7P sedoheptulose-7-phosphate

SBML Systems Biology Markup Language

SBP sedoheptulose-1,7-biphosphate

SBPase sedoheptulose bisphosphatase

SDH succinate dehydrogenase

Ser serine

SGT serine-glyoxylate aminotransferase

SHMT serine hydroxymethyl-transferase

STK succinate CoA ligase

Suc succinate

Abbreviations

IX

SucCoA succinyl-CoA

TAL transaldolase

TCA the tricarboxylic acid

THF tetrahydrofolate

Thr threnine

TKT transketolase

TPI triose phosphate isomerase

Tyr tyrosine

UDP uridine-diphosphate

UDPG UDP-α-D-glucose

UMP uridine monophosphate

UTP uridine-triphosphate

Val valine

Vc/Vo the flux ratio of carboxylation to oxygenation of RubisCO

Vi/Vo photon conversion efficiency

X5P xylulose-5-phosphate

фCO2 moles of CO2 fixed per mole of absorbed photons

X

Chapter 1 Introduction

Chapter 1

2

1.1 Plant metabolism and its investigation

Plants are indispensable for human life as they are the major source of food, feed

and fiber for human beings and livestock. Plant cells also serve as pharmaceutical

factories. Many of the drugs currently available on the market are plant-derived nat-

ural products or their derivatives (Rischer et al., 2013; Atanasov et al., 2015).

Additionally, some of the plant natural products are increasingly considered as

health-promoting compounds, such as lycopene in tomatoes and glucosinolates in

broccoli (Dixon, 2005; Hounsome et al., 2008).

The growth and survival of plants rely on metabolism which encompasses all

biochemical reactions mostly catalysed by enzymes within a cell (Smith and Stitt,

2007). Conventionally, metabolism is divided into two broad categories: catabolism

and anabolism. The former produces energy by breaking down organic molecules,

for example, the cellular respiration. The latter, anabolism synthesizes cellular com-

ponents such as sugars and proteins that are used for building the cell structure.

Biochemical reactions are organized into metabolic pathways which shape a

dynamic circuitry commonly referred as metabolic network (Sweetlove et al.,

2008). From a metabolic perspective, plants are highly versatile. They utilise light

energy to convert atmospheric CO2 directly into organic molecules required for life.

Compared to microbes, plant metabolism is extremely complex because of the sub-

cellular compartmentalization of metabolic processes and extensive secondary

metabolism. For instance, bacterial and yeast metabolisms contain only a few hun-

dred metabolites, while a commonly quoted estimate is that plants produce in the

order of 100,000 known plant secondary metabolites, including phenylpropanoids

and flavonoids, terpenoids, glucosinolates and alkaloids (Schwab, 2003). These sec-

ondary metabolites play important roles in cellular function, in signaling and in

adaptation to abiotic, for instance, drought, and biotic stress that cause damage to

plants via living organisms such as fungi, bacteria and insects. Moreover, plants

display enormous diversity in their metabolism in response to different environmen-

tal conditions and genetic conditions, which increases the complexity of plant

metabolism.

Cellular metabolism is regulated at different hierarchical levels (Figure 1.1).

While, fluxomics involves the quantitative analysis of metabolic fluxes through the

operative metabolic network, providing a holistic picture of integrated regulation at

the transcriptional, translational and metabolic levels. Metabolic fluxes, the rates of

metabolic reactions, are considered as ultimate representation of cellular phenotype

and function expressed under a certain condition (Nielsen, 2003). Metabolomics can

Introduction

3

also be useful to demonstrate metabolic signatures and gain a better understanding

on the interaction between plants and their environments (Field and Lake, 2011;

Kim et al., 2015), but typically fluxes are more informative than metabolite concen-

trations themselves, and it is challenging to estimate fluxes from metabolite

concentrations partly because each metabolite involves in multiple metabolic reac-

tions (Kim and Lun, 2014). Comprehensive characterization of plant metabolism

and their functions therefore requires quantitative knowledge of metabolic fluxes.

Over the past few years, much effort has been made to develop approaches for quan-

tifying metabolic fluxes, of which, metabolic flux analyses and constraint-based

modelling are the two most widely used techniques that take advantage of stoichio-

metric information of the metabolic network (Price et al., 2004).

Figure 1.1 Hierarchical layers involved in the final phenotype of an organism.

Metabolic flux analyses (MFA) is an experimental technique to quantify the

rates of material movement through complicated networks using isotopic labeling.

In this approach, the intracellular fluxes are calculated based on a stoichiometric

model for the major intracellular reactions and applying isotopic mass balances

Chapter 1

4

around intracellular metabolites. A set of fluxes that can be measured directly, typ-

ically the uptake rates of substrates into the cells and the rates of metabolites that

secreted from the cells (i.e. exchange fluxes), are used as input to do the calculations.

Various methods of MFA have been developed, the advantages and limitations of

each method have been discussed in the work of Antoniewicz (2015). MFA has been

applied to several plant studies and has been recognized as a valuable tool for ana-

lysing metabolic fluxes in plants (Kruger and Ratcliffe, 2009; Sweetlove and

Ratcliffe, 2011), however, the application of MFA to eukaryotic systems still re-

mains challenging due to a number of critical limitations such as the small size of

the networks studied (Sweetlove and Ratcliffe, 2011; Antoniewicz, 2015; Colombié

et al., 2015).

1.2 Constraint-based modelling

A complementary approach for quantifying metabolic flux is provided by in silico

constraint-based modeling (CBM), also known as structural metabolic modeling.

Constraint-based flux analysis rely on metabolic models that systematize biochem-

ical, genetic and genomic knowledge into a mathematic form, i.e. stoichiometry

matrix, and the steady-state mass balance assumption, does not require kinetic pa-

rameters. CBM enables the modelling of very large metabolic networks with a

minimum amount of input information and therefore has been the method of choice

for representing genome-scale reconstructed metabolic networks. In this thesis, the

constraint-based modeling approach was also used to study the metabolic behaviors

of tomato leaves. Therefore, here I introduce the basic concepts in the constraint-

based modelling and its combination with other techniques.

Stoichiometric matrix 𝑵, consists of rows corresponding to metabolites, col-

umns corresponding to reactions and each element 𝑵(𝑖, 𝑗) corresponding to the

stoichiometric coefficient of metabolite i in reaction j. The positive and negative

coefficients indicate the product and reactant metabolites, respectively. Stoichio-

metric matrix describes how the various reactions of the network are connected,

thus, is one of the most important predictors in the analysis of constraint-based mod-

els.

Steady-state assumption is the first principle of CBM to investigate the me-

tabolism of living cells, where net rate of any internal metabolite is zero. It is worth

noting that when modelling a metabolic system, metabolites are often classified into

two categories: internal metabolites and external metabolites. External metabolites

can be assumed as water or other species from the environment, such as photon and

Introduction

5

CO2, all the remaining intermediates are internal metabolites. The steady-state as-

sumption holds only for internal metabolites because each internal metabolite must

not accumulate or decrease in time. This assumption is supported by the observation

that intracellular dynamics are much faster than extracellular dynamics (Kauffman

et al., 2003). Therefore, it is sensible to disregard its transient behavior and consider

that they rapidly reach the steady state. The general equation for steady-state mass

balance assumption is described by:

𝑵𝒗 = 0

where 𝑵 is the stoichiometric matrix of the model and 𝒗 is a vector of metabolic

fluxes representing each feasible steady-state. As the number of reactions (columns)

is typically larger than the number of metabolites (rows) within a metabolic network

(Klamt et al., 2002), the system is thus underdetermined, resulting multiple possible

flux states that satisfy the steady-state assumption. To reduce the possible flux so-

lution space, additional constraints are generally imposed to the system, such as

information about the rates of substrate consumption and biomass synthesis.

1.2.1 Flux balance analysis

Optimization-based approaches seek to find the optimal solution that can be attained

by the cells, of which, flux balance analysis (FBA) is the most widely used one. It

uses linear programming (LP) to identify a specific (extreme) pattern of metabolic

fluxes based on the steady-state assumption and additional constraints such as reac-

tion capacities (Orth et al., 2010). The characteristic assumption of FBA is the

optimal function of the metabolic network termed as objective function, i.e. cells

regulate their fluxes toward optimal flux states. Therefore, the objective function

plays a crucial role in constraint-based flux analysis. The implemented objective

functions range from growth over ATP production and total flux minimization to

combined objectives. To evaluate the consequences of objective functions in FBA,

studies have been carried out to examine the choice of objective function optimiza-

tion with metabolic networks for predicting metabolic phenotypes (Schuetz et al.,

2007; Feist and Palsson, 2010; Cheung et al., 2013). In most cases, the biomass

objective function (growth rate or the rate of biomass production) that describes the

rate of which all of the biomass components (amino acids, nucleotides, etc.) are

made in their proportions has been widely used, in particular for micro-organisms.

The formulation of the biomass objective function for genome-scale network recon-

struction can be in previous studies (Becker et al., 2007; Feist and Palsson, 2010).

Chapter 1

6

The influence of biomass composition in the biomass objective function was exam-

ined in this thesis, using Arabidopsis models. The respective part is described in

Chapter 2.

1.2.2 Flux variability analysis

As FBA typically only provduces a unique steady-state flux distribution through a

metabolic network, and neglects the existences of multiple alternative flux solutions

exist that satisfy the optimization of a predefined objective function.Flux variability

analysis (FVA), an extension of FBA, has been proposed to determine all the possi-

ble flux ranges for each reaction flux that meet the optimal objective function

(Mahadevan and Schilling, 2003). Such alternative flux solutions refer to alternative

metabolic routes for achieving the same optimal behavior, and reveal the flexibility

and robustness of the underlying metabolic network.

1.2.3 Flux-sum analysis

At steady state, the net accumulation of internal metabolites is zero in the constraint-

based flux analysis, but the turnover rate, i.e. the overall consumption or production

rate of the intermediates, can be nonzero. To quantify the metabolite turnover rates,

flux-sum analysis (FSA) was proposed (Chung and Lee, 2009). The flux-sum of a

metabolite is calculated by halving the absolute sum of all incoming or outgoing

fluxes around the metabolite. The flux-sum for a metabolite can be used as a proxy

for the metabolite turnover and regarded as a measure for the overall flux through a

metabolic pool at a feasible steady state. Recently, it has been used to characterize

the variation in cellular metabolism under different conditions (Simons et al., 2014;

Basler et al., 2016).

1.3 Genome-scale metabolic models

The utility of constraint-based modelling in biological studies and the availability

of genome annotations allow the development of genome-scale metabolic networks.

It should be noted that there are two types of plant genome-scale metabolic networks

available: descriptive and predictive. The metabolic networks discussed in this the-

sis refer to the later, predictive genome-scale metabolic models (GSMs). The

descriptive genome-wide metabolic networks are not predictive and therefore have

limited capabilities that facilitate metabolic engineering, but they have been used

Introduction

7

for many studies (Bassel et al., 2012; Dharmawardhana et al., 2013; Kim et al.,

2015).

A GSM accounts for reaction stoichiometry and directionality, gene-protein-

reaction associations (GPR) associations, reaction compartmentalized localization

and transporter information as well as biomass composition. As an increasing num-

ber of the complete genome sequences of organisms become available, a large

number of GSMs have been constructed for several organisms in the domains of

bacteria, archaea and eukarya (Feist et al., 2009; Kim et al., 2012). GSMs have also

been successfully applied to plants, including Arabidopsis (Poolman et al., 2009; de

Oliveira Dal’Molin et al., 2010a; Cheung et al., 2013; Arnold and Nikoloski, 2014),

and food crops such as barley (Grafahrend-Belau et al., 2009; Grafahrend-Belau et

al., 2013) and rice (Poolman et al., 2013). Among these plant species, Arabidopsis

thaliana is the most extensively studied as Arabidopsis was the first plant to have

its genome sequenced (The Arabidopsis Genome Initiative, 2000). Moreover, Ara-

bidopsis serves as a model organism for research in plant biology and genetics

because of its important features, including a short generation time, small size, large

number of offspring and a relatively small genome. Since 2009, a handful of ge-

nome-scale/large-scale metabolic models for Arabidopsis have been published. In

this thesis, the published Arabidopsis GSMs were exploited, and the respective work

was described in Chapter 2.

Tomato (Solanum lycopersicum L.) is a high-value vegetable crop grown

worldwide in greenhouse as well as in open field. Tomato has always drawn much

attention owing to its high economic value in the market, and its high content in

health-promoting antioxidant compounds such as ascorbic acid (vitamin C) (Dixon,

2005; Story et al., 2010), as well as its distinct developmental features, for example,

compound leaves. In addition, tomato is widely used as an excellent model system

for addressing the fruit development and metabolism because of its ease of cultiva-

tion, and short life cycles (Klee and Giovannoni, 2011). The growth, yield and fruit

quality can be affected by several environmental factors, such as water deficit and

light, which induce the metabolic changes in cellular metabolism. Over the last few

decades, much effort has been made towards understanding the cellular behavior of

tomato, ranging from physiology and biochemistry to molecular and genetics by

conventional experimental techniques (Haupt-Herting et al., 2001; Lieberman et al.,

2004; Liu et al., 2004). However, the adaptive mechanisms of tomato under differ-

ent growth conditions are still not completely understood. The availability of tomato

genome allows systems-level studies of this commercial crop using constraint-based

approaches (Consortium, 2012). In this thesis, a genome-scale metabolic network

Chapter 1

8

of tomato was constructed, then was used to investigate metabolic behaviors under

drought and varying light conditions.

1.3.1 Reconstruction of GSMs

The reconstruction process of GSMs has been extensively reviewed (Feist et al.,

2009; Durot et al., 2009; Thiele and Palsson, 2010), of which, the most comprehen-

sive and detailed protocol was reported by Thiele and Palsson (2010). To

summarize, the steps toward a reconstruction of GSMs can be grouped into two

stages: initial reconstruction and model refinement (Figure 1.2).

Figure 1.2 Illustration of the GSMs’ reconstruction process. Adopted from (Durot et al., 2009).

Initial reconstruction. The reconstruction of a GSM initiates with the auto-

mated generation of a draft model based on the genome annotation of the target

organism. These genome information can be obtained from metabolic databases

such as KEGG (Kyoto Encyclopedia of Genes and Genomes; Kanehisa and Goto,

2000), BRENDA (Schomburg et al., 2004) and BioCyc (Karp et al., 2005). An au-

tomatically constructed metabolic network is, in essence, a collection of metabolic

reactions, always has gaps, errors and inconsistencies. Moreover, some information

Introduction

9

that is needed for the metabolic reconstruction such as subcellular localization, re-

action directionality and transporters, are not or falsely included in the metabolic

databases. Thus, the obtained draft model need to be manually curated in the next

stage.

Model refinement. In this stage, the obtained draft model is curated with re-

gard to the removal of metabolites and reactions, reaction stoichiometry, reaction

reversibility, mass conservation as well as energy consistency. The removed metab-

olites from the draft model are mainly non-metabolic and generic compounds.

Accordingly, reactions involved with ill-defined metabolites mentioned above were

removed. Reactions with wrong stoichiometry and reversibility were manually cor-

rected and replaced the correct ones based on literature study. Missing inter-

compartmental transport reactions were added based on an extensive literature

search.

Ensuring mass conservation is an important consideration during the develop-

ment of stoichiometric metabolic models. It is trivial to determine the atomic

balance of reactions for which the empirical formulas of all metabolites are known.

However, for those for which the empirical formula is not available must be treated

with some caution. Commonly, these reactions involve polymeric species such as

starch and cellulose. To solve this issue, the smallest monomeric subunit is defined,

and used for rescaling stoichiometric coefficients. For example, glucose is assumed

to be the monomeric subunit of starch and cellulose. Additionally, energy and redox

conservation was also tested during the refinement process. We ensured that ATP

and NAD(P)H could not be produced without nutrient inputs. The implement of

checking energy consistency has been detailed elsewhere (Thiele and Palsson, 2010;

Hartman et al., 2014).

Further in the model refinement process, the model is validated by examining

its ability of producing biomass metabolites, reproducing known metabolic behav-

iors and predicting growth phenotypes such as genotype-phonotype relationship.

Missing reactions and mistakenly included information can also be identified in

these analyses. In contrast to the rapid automated reconstruction step, the manual

curation process is very laborious as well as time-consuming, and it is usually an

iterative cycle. Previous publications, textbooks work are the main resources for the

manual curation step.

1.3.2 Software for genome-scale metabolic modelling

Due to the successful application of GSMs to investigate cellular metabolism, a

large number of software packages have been developed to aid and speed up the

Chapter 1

10

construction of new GSMs, thereby alleviating much of the manual effort previously

required. Features of these platforms have been extensively reviewed elsewhere

(Hamilton and Reed, 2014; Koussa et al., 2014; Lakshmanan et al., 2014; Martins

Conde et al., 2016). Here, I only discuss two software packages used in the thesis

work, which are Constraint-based reconstruction and analysis (COBRA; Becker et

al., 2007) and ScrumPy (Poolman, 2006).

COBRA methods are currently widely used in the field of genome-scale meta-

bolic modelling. The COBRA toolbox developed on MATLAB and also available

as a package for (Ebrahim et al., 2013) Python provides a high-level interface to a

variety of methods for constraint-based modelling of GSMs. Detailed documenta-

tions of COBRA are available at https://opencobra.github.io/. The structural

analyses of the Arabidopsis metabolic models in this study was performed using

COBRA toolbox.

ScrumPy is a metabolic modelling software package written in Python pro-

gramming language (Poolman, 2006). For detailed instructions of ScrumPy and its

usage, it is available from the main ScrumPy website

(http://mudshark.brookes.ac.uk/index.php/Software/ScrumPy). In this thesis work,

the construction of a genome-scale metabolic model of tomato and the subsequent

constraint-based analysis were carried out using ScrumPy.

1.3.3 Applications of GSMs

More and more GSMs have been reconstructed over the last decade (Kim et al.,

2012). These models have guided research in a variety of ways, including (1) con-

textualize high-throughput data (Lakshmanan et al., 2015; Mohanty et al., 2016),

(2) guide metabolic engineering (Basler et al., 2016), (3) direct hypothesis-driven

discovery, (4) interrogate multispecies relationships (de Oliveira Dal’Molin et al.,

2015), (5) analyze biological network properties (Oberhardt et al., 2009; McCloskey

et al., 2013) and (6) predict cellular phenotypes (Lakshmanan et al., 2013; Simons

et al., 2014; Seaver et al., 2015). In addition, GSMs allow integrating additional

data such as transcriptomics and proteomics data sets, which together with bioinfor-

matics approaches can support a better understanding of metabolic behavior in

plants (Töpfer et al., 2013). Altogether, these studies have established mechanistic

understandings of genotype–phenotype relationships and metabolic characteristics

of cellular metabolism, at a genome-scale level.

Introduction

11

1.4 Organization of this thesis

As mentioned earlier, the growth, yield and fruit quality of tomato can be affected

by several environmental factors. To better understand the cellular behavior under

drought and varying light conditions, this thesis aims to develop an in silico model

to characterize tomato’s metabolic behavior. In addition, the impact of the model-

ling parameters such as the biomass compositions and maintenance in flux-balance

modelling of plant metabolism was assessed. The remainder of this thesis is orga-

nized as follows:

Chapter 2 describes how flux solutions predicted by FBA are sensitive to the

biomass composition and model structure. The analyses are applied to Arabidopsis

which is the best-investigated plant model organism. We found that the structure of

the metabolic model has a large impact on FBA predictions in comparison with the

biomass composition. This chapter also provides a succinate review of the existing

plant GSMs.

Chapter 3 describes the reconstruction of the first genome-scale metabolic

model of tomato using ScrumPy. The model is then tested by comparing the activity

of well characterized pathways in silico with biochemical evidence in the literature.

In Chapter 4, the reconstructed tomato model described in Chapter 3 is applied

to characterize the metabolic behaviors of leaves for photorespiration and other rel-

evant metabolic processes under drought conditions. The essential

reactions/enzymes under both normal and drought conditions are also presented.

Chapter 5 describes the extension of the tomato model to a light-specific model

by incorporating the light-driven photophosphorylation reactions in a wavelength-

specific manner. The model is then used to investigate how changes in light intensity

and quality affect cellular growth and how the energy ATP and the reducing equavi-

lant NADPH is regulated accordingly. Moreover, the model is applied to evaluate

the light source efficiency.

Chapter 6 summarizes the main contributions made in this thesis and discusses

potential extensions of the present work.

Chapter 2 Flux balance analysis of

plant metabolism: the effect of

biomass composition and model

structure on model predictions1

1 This chapter has been published:

Yuan, H., Cheung, C.Y.M., Hilbers, P.A.J. and van Riel, N.A.W. (2016) Flux

balance analysis of plant metabolism: the effect of biomass composition and

model structure on model predictions. Front. Plant Sci.7, 537.

Abstract

The biomass composition represented in constraint-based metabolic models is a key

component for predicting cellular metabolism using flux balance analysis (FBA).

Despite major advances in analytical technologies, it is often challenging to obtain

a detailed composition of all major biomass components experimentally. Studies

examining the influence of the biomass composition on the predictions of metabolic

models have so far mostly been done on models of microorganisms. Little is known

about the impact of varying biomass composition on flux prediction in FBA models

of plants, whose metabolism is very versatile and complex because of the presence

of multiple subcellular compartments. Also, the published metabolic models of

plants differ in size and complexity. In this chapter, we examined the sensitivity of

the predicted fluxes of plant metabolic models to biomass composition and model

structure. These questions were addressed by evaluating the sensitivity of predic-

tions of growth rates and central carbon metabolic fluxes to varying biomass

compositions in three different genome-/large-scale metabolic models of Arabidop-

sis thaliana. Our results showed that fluxes through the central carbon metabolism

were robust to changes in biomass composition. Nevertheless, comparisons between

the predictions from three models using identical modelling constraints and objec-

tive function showed that model predictions were sensitive to the structure of the

models, highlighting large discrepancies between the published models.

Sensitivity in plant metabolic models

15

2.1 Introduction

Flux balance analysis (FBA), a constraint-based modelling approach, is widely used

in predicting metabolic fluxes based on stoichiometric metabolic models, in partic-

ular, large-scale or genome-scale metabolic models (GSMs) (Orth et al., 2010).

Stoichiometric metabolic models are typically underdetermined because the number

of reactions in the model is usually larger than the number of metabolites (Bonarius

et al., 1997; Kauffman et al., 2003). Therefore, in most cases, constraint-based anal-

ysis yields multiple feasible flux solutions. To narrow down the space of feasible

solutions, additional constraints can be imposed by specifying the range of fluxes

through any particular reaction. In addition to the application of constraints, an ob-

jective function is usually defined for identifying biologically relevant flux

solutions. The most commonly used objective function for FBA is the biomass ob-

jective function (BOF), which is to maximize the efficiency of biomass production,

i.e. growth rate (Feist and Palsson, 2010). Biomass production is mathematically

represented by a so called ‘biomass reaction’ which, in essence, is a collection of all

individual biomass constituents together with their fractional contributions to the

overall cellular biomass, and energetic requirements for the biomass generation.

Knowledge of the biomass composition is crucial for predicting flux distribu-

tion in metabolic models using FBA because the intracellular fluxes are dependent

on the fluxes contributing to biomass synthesis (Pramanik and Keasling, 1997;

Schwender and Hay, 2012). Therefore, an important consideration during the devel-

opment of GSMs is to define the biomass composition, ideally for the condition

under study. Experimental evidence indicates that the biomass composition varies

between species, cell types and physiological conditions (Novák and Loubiere,

2000; Hay and Schwender, 2011). However, due to a lack of organism-specific

and/or condition-specific experimental information, the biomass compositions used

in plant GSMs are often collected from diverse types of measurements, experiments,

research groups, and even different cell types and plant species (Collakova et al.,

2012).

Some computational methods have been developed to estimate the fractional

contribution of a precursor to the biomass reaction in microorganisms, e.g. calculat-

ing the coefficients of deoxy-nucleotide triphosphates (dNTPs) and nucleotide

triphosphates (NTPs) according to the fraction of DNA and RNA (Orth et al., 2010;

Thiele and Palsson, 2010). Nevertheless, these approaches can only be used for

amino acids, NTPs (ATP, GTP, CTP, UTP), and dNTPs (dATP, dGTP, dCTP,

Chapter 2

16

dTTP). Given the existence of multiple organelles in plants, one needs to be cautious

when applying these approaches to plant models.

With the increasing use of network reconstructions and constraint-based ap-

proaches, a need has arisen to clearly define and demonstrate the relevance of the

modeling parameters, such as biomass composition, for predicting metabolic fluxes.

Work examining the influence of the biomass composition on the predicted fluxes

has mostly been done on models of microorganisms, in particular E. coli (Pramanik

and Keasling, 1997; Feist et al., 2007). Most recently, a sensitivity analysis of a

yeast model suggested that model predictions are sensitive to variations in biomass

composition (Dikicioglu et al., 2015). These observations naturally lead to questions

about the sensitivity of flux predictions in plant metabolic networks to biomass com-

position and the robustness of plant metabolic models. Thus far, there are very

limited studies exploring the effects of changes in biomass composition on the flux

distributions in plants. Plant metabolic networks are significantly more complex

than those of microorganisms due to the presence of multiple compartments and

parallel metabolic pathways. A study on a model of oilseed rape suggested that flux

predictions are sensitive to the contents of oil and protein, which are the major stor-

age components in oil seed (Schwender and Hay, 2012). However, in a study of

Arabidopsis heterotrophic cell culture, central carbon metabolism has been ob-

served to be robust to different conditions despite the significant differences in the

resulting biomass compositions (Williams et al., 2010). Given that plants are

adapted to grow in diverse environmental conditions, plant metabolism is expected

to be flexible in face of perturbations. Thus, it deserves theoretical exploration on

basis of constraint-based metabolic models to assess the influence of changing bio-

mass composition on predicted fluxes.

Arabidopsis, a model organism for plant biology, has been studied extensively

with systems-biology approaches. In this study, we started by reviewing the pub-

lished Arabidopsis metabolic models followed by an investigation of the impact of

changing the biomass composition on the flux predictions in large- or genome-scale

plant metabolic models, in particular, the fluxes through central metabolic pathways

(i.e. glycolysis, pentose phosphate pathway (PPP), TCA cycle, and mitochondrial

electron transport chain (ETC)). This is because these existing large-scale metabolic

networks of plants provided mostly qualitative predictions of intracellular fluxes for

primarily central carbon metabolism. Furthermore, previous work has shown that

fluxes of central carbon metabolism dominate the FBA results, with little to no flux

through the secondary metabolic pathways (Collakova et al., 2012). In this chapter,

we focused on study on three published models of Arabidopsis, which have different

Sensitivity in plant metabolic models

17

biomass compositions and network structures. We systematically evaluated the in-

fluence of biomass composition on the outcome of FBA simulations in three ways:

(1) using different biomass compositions with the same model; (2) using the same

biomass composition with different models; (3) varying individual components of

the biomass composition and maintenance cost. Our analyses indicate that i) the

central metabolic fluxes are relatively stable in face of varying biomass composi-

tion, regardless of model structure; and ii) the model structure is the main factor in

determining the variation in computational results generated by using FBA.

2.2 Methods

2.2.1 Stoichiometric models

In this study, we compared and investigated three published stoichiometric models

of Arabidopsis, denoted as Poolman (Poolman et al., 2009), AraGEM (de Oliveira

Dal’Molin et al., 2010a) and AraCore (Arnold and Nikoloski, 2014). The Systems

Biology Markup Language (SBML) format for the Poolman and AraCore model

were available from supplementary files of the corresponding paper. The direction

of the phenylpyruvate carboxylase reaction in the Poolman model has been cor-

rected as reported in their subsequent publication (Williams et al., 2010). For the

AraGEM model, an updated version was obtained from

http://web.aibn.uq.edu.au/cssb/resources/Genomes.html. In this study, we simu-

lated the cellular metabolism of Arabidopsis cells growing on glucose as carbon and

energy sources under aerobic heterotrophic conditions.

Although all major biomass components (i.e. cell wall, protein, lipid, carbohy-

drate, DNA and RNA) were taken into account in the three models via representative

metabolites or corresponding precursors, the biomass components included in these

three models are not exactly the same. For example, xylose and γ-aminobutyric acid

(GABA) were not considered in Poolman model, but were included in AraGEM and

AraCore models. Similarly, soluble metabolites were only considered in AraCore

model, but not in Poolman and AraGEM models. To keep the list of biomass com-

ponents as consistent as possible between the three models, we added some

additional biomass transporters to the models if required. It is noted that not all

newly added biomass components can be produced by all models, e.g. maltose, xy-

lose and GABA cannot be produced in Poolman model, coniferyl-alcohol,

coumaryl-alchol, sinapyl-alcohol and xylose cannot be produced by AraCore model

as these metabolites are not present in AraCore. Therefore, in this study, we only

Chapter 2

18

added new biomass components that can be produced by all the three Arabidopsis

models. In total, we added 9 new biomass transporters to Poolman model and 7 for

AraGEM (Appendix 2.6.1).

2.2.2 Biomass equations

The biomass compositions used in this study were extracted from previous studies,

hereafter referred to as PoolmanBOF, AraGEMBOF, and AraCoreBOF, corre-

sponding to the biomass composition used in Poolman, AraGEM and AraCore

model, respectively (Appendix 2.6.2). Since three condition-specific biomass com-

positions (reflecting carbon-limiting, nitrogen-limiting and optimal growth

conditions) have been employed in the AraCore model, here we chose the carbon-

limiting biomass reaction to represent AraCoreBOF. The original biomass compo-

sition of each model, together with the calculation of weight percentage of biomass

components (Figure 2.1) is provided in Appendix 2.6.2.

Due to the different units originally used in the three studies (mmol/g/L in

Poolman and mmol/g DW in AraGEM and AraCore), the biomass compositions

were normalized to enable a fair comparison. These calculations were performed by

weight, defining 1 unit of flux through the biomass equation equals to 1g of biomass.

The details of these calculations are provided in Appendix 2.6.2. Finally, we used

the normalized biomass equations to perform our model simulations.

2.2.3 Model simulations

FBA was used to determine the flux solutions at steady-state condition. All simula-

tions were performed using FBA in geometric mode (Smallbone and Simeonidis,

2009) as implemented in the COBRA toolbox (Becker et al., 2007) executed in

MATLAB (The MathWorks, version R2012a). Geometric FBA enables a unique

optimal solution that is central to the range of possible flux distributions. The Gurobi

Optimizer (http://www.gurobi.com, version 5.0.2) solver in combination with the

COBRA toolbox were used to solve the linear programming problems. In our study,

the biomass equation was maximized to obtain the optimal solution of the metabolic

model as described elsewhere (Orth et al., 2010a). Formally, the FBA problem can

be stated as follows:

Sensitivity in plant metabolic models

19

Maximize: 𝒗𝒈𝒓𝒐𝒘𝒕𝒉

GAM 1i

growth

ii

vBiomassc X

Subject to 𝑺𝒗 = 0

and 𝒗𝒎𝒊𝒏 ≤ 𝒗 ≤ 𝒗𝒎𝒂𝒙

where 𝒗𝒈𝒓𝒐𝒘𝒕𝒉 is the flux that the biomass reaction carries, representing growth rate,

𝑐 is the vector of biomass coefficients, whose component ci indicates the ratio of

metabolite Xi required for the formation of a unit of biomass, 𝑺 is the stoichiometric

matrix, 𝒗 is a vector of all reaction fluxes in the system, also referred to as the flux

distribution, 𝒗𝒎𝒊𝒏 and 𝒗𝒎𝒂𝒙 represent lower and upper bounds for the flux of each

reaction, respectively. GAM refers to growth associated maintenance.

Figure 2.1 Weight percentage of the biomass components. The original and modified (used for simulations) weight

percentage for each class of metabolites contributing to biomass synthesis is displayed. The composition is dis-

played for Poolman model (a), AraGEM model (b), and AraCore model (c). The calculations for each class of

metabolites are shown in Appendix 2.6.2.

To simulate the cellular behavior of Arabidopsis cells, we constrained the glu-

cose uptake rates at 10 flux units, which was the only source of carbon and energy.

AraCore represents a photoautotrophic cell, which does not have organic sources

for heterotrophic scenarios, but it can also be utilized to simulate heterotrophic con-

ditions by adapting the energy source (Arnold and Nikoloski, 2014). Consequently,

Chapter 2

20

we added an additional glucose exchange reaction ‘Im_Glc’ to AraCore model. Am-

monia (NH3) and hydrogen sulfide (H2S) were constrained to be utilized as the sole

N and S sources, respectively, because AraGEM and AraCore can only grow with

H2S and NH3 as the S and N sources, respectively. Since SO42- is the sole S source

in the Poolman model, it does not have an H2S transporter. We therefore added an

H2S transporter to Poolman model, namely ‘H2S_tx’, enabling H2S as sole sulfur

source in all models. Pi is the sole P source input in all three models. Additionally,

non-growth associated maintenance (NGAM) was included in all simulations with

a value of 2.02, which is a normalized flux unit referring to the value reported in

Poolman model (Appendix 2.6.2). Similarly, GAM was fixed at 53.26 flux units in

this study, which was scaled based on the value reported in AraGEM model. Im-

ported or exported metabolites are always freely exchangeable across the system

boundary to provide the necessary nutrients and remove secreted substances.

The biomass equation (i.e. BOF) is generated by defining all of the biomass

constituents, in which all the precursor metabolites are assembled in one single re-

action with corresponding coefficients (Feist and Palsson, 2010; Thiele and Palsson,

2010). For a fair comparison, we only considered the biomass components which

can be produced by all three models in the three biomass equations (Appendix

2.6.2). This results in new biomass compositions for AraGEM and AraCore models

because some biomass metabolites such as xylose and maltose that cannot be pro-

duced by all 3 models are not taken into account in our biomass reaction. The

modified weight percentages of biomass compositions that are included in our bio-

mass reactions differs slightly from the original ones for the three models (Figure

2.1).

To investigate the influence of biomass composition on the predicted fluxes,

we performed three scenarios in each model with the biomass equations of

PoolmanBOF, AraGEMBOF and AraCoreBOF, respectively. We replaced the co-

efficient of each biomass component in the biomass equation with the corresponding

values in the other two models. In total, 9 scenarios were simulated in the study,

namely ‘Poolman-PoolmanBOF’, ‘Poolman-AraGEMBOF’, ‘Poolman-AraCore-

BOF’, ‘AraGEM-PoolmanBOF’, ‘AraGEM-AraGEMBOF’, ‘AraGEM-

AraCoreBOF’, ‘AraCore-PoolmanBOF’, ‘AraCore-AraGEMBOF’, and ‘AraCore-

AraCoreBOF’. To assess the differences between predicted fluxes obtained from

different scenarios, we define the biomass composition given in each model as the

‘reference’ scenario. Thus, ‘Poolman-PoolmanBOF’, ‘AraGEM-AraGEMBOF’,

and ‘AraCore-AraCoreBOF’ are the ‘reference’ scenarios in Poolman, AraGEM,

and AraCore model, respectively.

Sensitivity in plant metabolic models

21

To confirm the confidence of the predictions, flux variability analysis (FVA)

was conducted for the 9 model-biomass combinations, which determines the range

of possible solutions for each reaction while giving rise to the same optimal value

for the objective function (Mahadevan and Schilling, 2003).

2.3 Results

2.3.1 Comparisons between plant flux-balanced models

Origins and uses of biomass compositions in plant flux-balanced models

To understand how the biomass composition data is obtained in published metabolic

models, we surveyed the source of data used in formulating biomass equations in

the existing large-scale metabolic models of plants (Table 2.1). From the survey, it

was verified that only 5 out of 21 models had their biomass compositions measured

by the research group that constructed the model, whereas, the remaining 16 studies

either do not include any biomass information or adopted from other research

groups, some of which were from other organisms. Furthermore, within the sur-

veyed plant metabolic networks, 6 of 21 (29%) used biomass data as the objective

function, 11 of 21 (52%) used biomass data as constraints, whereas the rest did not

perform any FBA simulations.

General properties of published Arabidopsis flux-balanced models

To investigate the variability of the biomass compositions and structure of the mod-

els for the same species, we chose Arabidopsis since eight metabolic models were

published since 2009. The general statistics of the available Arabidopsis models to

date is summarised in Table 2.2. In general, we observed an increase in the number

of genes, metabolites, reactions and transporters included in Arabidopsis metabolic

models over time, but this increase was not uniform. For example, the number of

metabolites and reactions in the model of Mintz-Oron et al. (2012) is much larger

than that of AraGEM (de Oliveira Dal’Molin et al., 2010a), but the latter has more

genes than the former. Towards the compartmentalization, the organelles included

in each model are relatively invariant over time. The various Arabidopsis genome-

scale models were reviewed more extensively elsewhere (Collakova et al., 2012; de

Oliveira Dal’Molin and Nielsen, 2013; Baghalian et al., 2014; Arnold et al., 2015).

Chapter 2

22

Table 2.1 Summary of the origin of biomass data in the existing large-scale metabolic models of plants.

Besides the general statistics, these models differ in several aspects, for in-

stance, the inclusion of cellular maintenance costs (Table 2.3). The models are quite

different regarding cell maintenance, because its experimental quantification is a

major challenge (Sweetlove et al., 2013).

Species No. Model Biomass composition data Biomass

used as

Arabidopsis

(Arabidopsis

thaliana)

1 Poolman et al.,

2009 Experimental Constraint

2

de Oliveira

Dal’Molin et al.,

2010a

Literature Guinn, 1966; Poorter and Bergkotte, 1992;

Niemann et al., 1995) Constraint

3 Radrich et al., 2010 No biomass No

simulation

4 Saha et al., 2011

Literature (Spector, 1956; Muller et al., 1970; Pen-

ning de Vries et al., 1974 ; Wedig et al., 1987) or

other related organisms

No

simulation

5 Mintz-Oron et al.,

2012

Literature (Weise et al., 2000; Reinders et al., 2005;

Poolman et al., 2009; de Oliveira Dal’Molin et al.,

2010a)

No

simulation

6 Chung et al., 2013 No biomass No

simulation

7 Cheung et al., 2013 Experimental Constraint

8 Arnold and

Nikoloski, 2014

Literature (Döermann et al., 1995; Sharrock and

Clack, 2002; Mooney et al., 2006; DeBolt et al.,

2009; Tschoep et al., 2009; Pyl et al., 2012; Sulpice

et al., 2013)

Objective

Barley

(Hordeum vul-

gare)

9 Grafahrend-Belau

et al., 2009 Literature (OECD, 2004) Objective

10 Grafahrend-Belau

et al., 2013

Literature (Antongiovanni and Sargentini, 1991;

Bonnett and Incoll, 1993a,b)

Constraint

Rapeseed

(Brassica

napus)

11 Hay and

Schwender, 2011

Biomass macromolecules determined experimen-

tally, the composition of biomass macromolecules

obtained from literature (Katterman and Ergle, 1966;

Norton, 1989; Schwender and Ohlrogge, 2002; Town

et al., 2006)

Constraint

12 Pilalis et al., 2011 Literature (Schwender et al., 2004; Schwender et al.,

2006) Objective

Maize

(Zea mays)

13

de Oliveira

Dal’Molin et al.,

2010b

Literature (Guinn 1966; Poorter et al.,1992; Niemann

et al., 1995) Constraint

14 Saha et al., 2011

Literature (Spector, 1956; Muller et al., 1970; Pen-

ningd et al., 1974; Wedig et al., 1987) or other

related organisms

Objective

15 Simons et al., 2014 Experimental Objective

Sorghum

(Sorghum

bicolor)

16

de Oliveira

Dal’Molin et al.,

2010b

Literature (Guinn, 1966; Poorter and Bergkotte,

1992; Niemann et al., 1995) Constraint

Sugarcane

(Saccharum

officinarum)

17

de Oliveira

Dal’Molin et al.,

2010b

Literature (Guinn, 1966; Poorter and Bergkotte,

1992) Constraint

Rice

(Oryza

sativa)

18 Poolman et al.,

2013 Literature (Juliano, 1985; Kwon and Soh, 1985) Constraint

19 Lakshmanan et al.,

2013 Literature (Juliano, 1985; Edwards et al., 2012) Objective

Tomato

(Solanum

lycopersicum)

20 Colombié et al.,

2015 Experimental Constraint

21 Yuan et al., 2016

Literature (Sheen, 1983; Roessner-Tunali et al.,

2003; Schauer et al., 2005; Nunes-Nesi et al., 2007;

Sánchez-Rodríguez et al., 2010; EI-Sayed, 2013)

Constraint

Sensitivity in plant metabolic models

23

Three Arabidopsis models have different biomass compositions

Poolman, AraGEM and AraCore models each use a different biomass composition

to simulate cell growth. Although the relative amounts of each biomass component

for all three models were derived from experimental data, they relied entirely on

different sources. Poolman model used measurements from their own group for

modelling a heterotrophic cell culture, while, AraGEM and AraCore used data for

various tissues for Arabidopsis or related species. As a further comparison, we ana-

lysed the macromolecular compositions for each model (Appendix 2.6.2), finding

the three models differ significantly in the composition of biomass macromolecules

(Figure 2.1). In Poolman model, cell wall comprised more than half of cell biomass.

In contrast, cells contain higher amounts of protein in AraGEM and AraCore model.

Experimental evidence indicates that the distributions of biomass components are

very tissue-specific (Mueller et al., 2003). One would expect that the biomass com-

positions of AraGEM and AraCore to be similar as both models represent

photosynthetic leaf cells. While there are some similarities in terms of the propor-

tions of cell wall, carbohydrate and protein, there are also major differences. For

example, the proportion of lipid in AraCore biomass is much larger than in AraGEM

biomass (18.4% versus 1.1%). This could be explained by the fact that biomass data

in AraGEM model is collected from different organisms.

Differences in central metabolism between the three Arabidopsis models

Going beyond merely comparing general network characteristics, we compared the

models at the individual reaction level, concentrating on the compartmentation and

reversibility of reactions in central carbon metabolism, which is an essential biolog-

ical process to sustain growth and biomass synthesis (Appendix 2.6.3). We focused

on three and biomass synthesis (Appendix 2.6.3). We focused on three Arabidopsis

models, Poolman, AraGEM and AraCore models, which have distinct model struc-

tures and biomass compositions. Generally, the three models covered all the listed

central metabolic reactions (Appendix 2.6.4). Compared with the other two models,

AraGEM did not include PPi-dependent phosphofructokinase (EC 2.7.1.90),

Chapter 2

24

Table 2.2 Structure comparison of the reconstructed metabolic models for Arabidopsis.

Symbol ‘-‘indicates information is unknown; ‘a’ indicates information is extracted from the original models, ‘b’ indicates information is retrieved from papers. Exchange reactions

allow for exchange of specific metabolites with the extracellular space. Transporters indicate metabolites can move between intracellular organelles. ‘c’, cytosol; ‘m’, mitochon-

dria; ‘p’, plastid; ‘x’, peroxisome; ‘v’, vacuole; ‘g’, Golgi; ‘e’, endoplasmic reticulum.

Reference Abbreviation Year Organelles Number of

genes

Number of

metabolites

Number of

reactions

Number of

exchange reactions

Number of intracellular

transporters

Poolman et al. (2009) Poolman[a] 2009 2

(c,m) Not available 1,253 1,406 42 -

de Oliveira Dal’Molin et al. (2010) AraGEM[a] 2010 5

(c,m,p,x,v) 1,419 1,737 1,601 18 81

Radrich et al. (2010) Radrich[a] 2010 - 1,571 2,328 2,315 - -

Saha et al. (2011) iRS1597[a] 2011 5

(c,m,p,x,v) 1,597 1,820 1,844 18 81

Mintz-Oron et al. (2012) Mintz-Oron[a] 2012 7

(c,m,p,x,v,g,e) 1,223 2,930 3,508 101 772

Chung et al. (2013) iAT1475[b] 2013 4

(c,m,p,x) 1,475 1,761 1,895 22 86

Cheung et al. (2013) Cheung[a] 2013 5

(c,m,p,x,v) 2,857 2,739 2,769 20 192

Arnold and Nikoloski, 2014 AraCore[a] 2014 4

(c,m,p,x) 634 407 549 98 124

Sensitivity analysis of plant metabolic models

25

Table 2.3 Network characteristics and FBA simulations in Arabidopsis GSMs.

‘PoolmanBOF’ indicates the biomass objective function included in the Poolman model; ‘AraGEMBOF’ indicates

the biomass objective function included in the AraGEM model; ‘AraCoreBOF’ indicates the biomass objective

function included in the AraCore model. GAM, growth associated maintenance; NGAM, non-growth associated

maintenance. [a]5 objective functions are minimization of overall flux, maximization of biomass, minimization of

glucose consumption, maximization of ATP production and maximization of NADPH production.

NADP-dependent non-phosphorylating glyceraldehyde-3-phosphate dehydrogen-

ase (EC 1.2.1.9) or NAD-dependent 6-phosphogluconate dehydrogenase (EC

1.1.1.343). Moreover, AraGEM uses two lumped reactions to represent the electron

transport chain (ETC) reactions, which are referred to as alternative oxidase path-

way (AOX) and cytochrome C oxidase pathway (COX). In contrast, Poolman and

AraCore models incorporate separate reactions to describe the ETC. Beyond these

differences, the localization and directionality of central metabolic reactions in the

three models are not always consistent. For example, the reaction catalyzed by phos-

phoenolpyruvate carboxylase (EC 4.1.1.31) in Poolman model operates in the

opposite direction compared to other two models, which was corrected in a subse-

quent publication (Williams et al., 2010). For AraCore, some reactions known to be

reversible, such as aconitase (EC 4.2.1.3) and fumarase (EC 4.2.1.2), are set as irre-

versible. Poolman model is not fully compartmentalized, so most of the reactions

involved in central carbon metabolic pathways are assigned to the cytosol.

Items

Model Cell type

Model

format

Cell

maintenance

Objective

function

BOF

included

Poolman Heterotrophic ScrumPy &

SBML NGAM Minimize total flux PoolmanBOF

AraGEM Photosynthetic &

Heterotrophic SBML GAM

Minimize photon/su-

crose uptake of

growth rate

AraGEMBOF

Radrich Unknown SBML Not included Not included Not included

iRS1597 Photosynthetic &

Heterotrophic Excel GAM Maximize biomass AraGEMBOF

Mintz-Oron Photosynthetic &

Heterotrophic SBML GAM

Minimize metabolic

adjustment (MOMA) AraGEMBOF

iAT1475 Photosynthetic &

Heterotrophic Excel GAM

Maximize IPP

production AraGEMBOF

Cheung Photosynthetic &

Heterotrophic

ScrumPy &

SBML GAM & NGAM

5 objective

functions[a]

Biomass as

constraints

AraCore Photosynthetic &

Heterotrophic SBML Not included

Maximize biomass &

energy efficiency AraCoreBOF

Chapter 2

26

2.3.2 The impact of biomass composition and model structure on

central metabolic fluxes

Central carbon metabolism is robust to changes in biomass compositions

To test the sensitivity of FBA solutions to the biomass composition, nine scenarios

described in the Methods section were simulated with three flux-balanced metabolic

models of Arabidopsis, Poolman, AraGEM and AraCore models, in combination

with their respective biomass compositions. The maximization of growth rate is

used as the objective function subjecting to mass-balance constraints, and setting

the glucose uptake rates as 10 flux units (section 2.2.3 Model simulations). In par-

ticular, we analysed how central metabolic reactions listed in Appendix 2.6.4

respond to a change in biomass composition. The results are illustrated in Figure 2

in which the differences in color intensity between columns (compare horizontally)

reflects the differences in flux values of each reaction calculated by the three bio-

mass compositions. It can be seen that for the majority of the reactions, in particular

the glycolytic reactions, the flux patterns were very similar. This indicates a high

stability of the central carbon metabolism in Arabidopsis with respect to biomass

composition, regardless of model structures.

Some reactions showed larger flexibility in our predictions. The flux distribu-

tion predicted by AraCore model with AraGEMBOF resulted in flux through 2-

oxoglutarate dehydrogenase reaction (reaction 17), a reaction belonging to the TCA

cycle, whereas this reaction did not carry flux with PoolmanBOF and AraCoreBOF.

The flux distributions through the TCA cycle reflect the function of the metabolic

network, and its operation largely depends on the cell type and the considered phys-

iological context (Sweetlove et al., 2010). In addition, 2-oxoglutarate (2-OG) is an

essential intermediate for the biosynthesis of amino acids such as glutamine and

glutamate (Appendix 2.6.3). Given that AraGEMBOF contains a much higher

amount of glutamine and glutamate (Appendix 2.6.2), itit is not surprising that 2-

oxoglutarate dehydrogenase reaction carried non-zero flux implemented with Ar-

aGEMBOF in AraCore model. For AraGEM model, the biomass composition

greatly affected the fluxes through glyceraldehyde-3-phosphate dehydrogenase (EC

1.2.1.9/1.2.1.12; reaction 7 and 8), phosphoglyate kinase (EC 2.7.2.3; reaction 9),

phosphoglycerate hydratase (EC 4.2.1.11; reaction 11), pyruvate kinase (EC

2.7.1.40; reaction 12) and malate dehydrogenase (EC 1.1.1.37; reaction 21). For ex-

ample, the use of PoolmanBOF on the AraGEM model gave rise to higher fluxes

through malate dehydrogenase compared to using AraGEMBOF, which is due to a

Sensitivity in plant metabolic models

27

Figure 2.2 Flux maps of central carbon metabolism predicted from three Arabidopsis models: Poolman model,

AraGEM model and AraCore model. Fluxes were predicted using three different biomass compositions in each

model: PoolmanBOF, the biomass composition included in Poolman model; AraGEMBOF, biomass composition

included in AraGEM model; AraCoreBOF, biomass composition included in AraCore model. Each reaction is

numbered, referencing Appendix 2.6.4, and the color intensity of each box corresponds to the flux value (mmol g-

1 DW hr-1) for the respective labeled reaction in each scenario. The results calculated by different biomass compo-

sitions with the same model can be interpreted by comparing between columns (compare horizontally). The results

calculated by the same biomass composition with different models can be interpreted by comparing between rows

(compare vertically). DW, Dry cell weight. Metabolite abbreviations are as follows: GLC, glucose; 2-OG, 2-ox-

oglutarate; 3-PGA, 3-phosphoglycerate; 2-PG, 2-phosphoglycolate; G6P, glucose-6-phosphate; F6P, fructose-6-

phosphate; 6PGL, 6-phosphogluconolactone; 6PG, 6-phosphogluconate; Ru5P, ribulose-5-phosphate; R5P, ribose-

5-phosphate; X5P, xylulose-5-phosphate; S7P, sedoheptulose-7-phosphate; E4P, erythrose-4-phosphate; FBP,

fructose-1,6-biphosphate; DHAP, dihydroxyacetone phosphate; DPG, glycerate-1,3-bisphosphate; PEP, phosphoe-

nol pyruvate; OAA, oxalacetic acid; GAP, glyceraldehyde-3-phosphate; 2-PGA, 2-phosphoglycerate; Pyr,

pyruvate; Cit, citrate; IsoCit, threo-isocitrate; Suc, succinate; SucCoA, succinyl-CoA; Fum, fumarate; Mal, malate;

QH2, ubiquinone; Q, ubiquinol; Cytred, cytochrome reduced; Cytox, cytochrome oxidized; Fdred, ferredoxin reduced;

Fdox, ferredoxin oxidized.

Chapter 2

28

reaction cycle that malate dehydrogenase involves in with PoolmanBOF. Com-

monly, the great changes occurred at the branch points of glycolysis and the TCA

cycle, where there were drains for the synthesis of cellular constituents. Under het-

erotrophic conditions, it is generally thought that the oxidative PPP (OPPP)

predominately provides reducing power for the production of biomass, in particular

for fatty acid synthesis. As a result, we would expect the OPPP reactions to be active

in our simulations. However, OPPP reactions carried no flux in any of the predicted

solutions of Poolman and AraGEM model. Interestingly, for AraCore model, using

AraGEMBOF resulted in small flux through OPPP reactions such as glucose 6-

phosphate (EC 1.1.1.49; reaction 22), 6-phosphogluconolactonase (EC 3.1.1.31; re-

action 23) and 6-phosphogluconate dehydrogenase (EC 1.1.1.44/1.1.1.343; reaction

24, 25). Previously, Williams et al. (2010) showed that the OPPP was poorly pre-

dicted by Poolman model, which was suggested to be caused by the error in

assigning the reversibility of the NADP+-dependent glyceraldehyde-3-phosphate

dehydrogenase (EC 1.2.1.13) (Cheung et al., 2013).

Model structure has a large impact on model predictions

Furthermore, we assessed the effects of model structures on the flux predictions by

comparing the model predictions from the same biomass equation with different

models (Figure 2.2, comparisons between rows; compare vertically). It is apparent

that the vast majority of central metabolic reactions carried different fluxes from

each other, despite the application of identical biomass composition and boundary

constraints. For example, the hexokinase reaction (reaction 1), which converts glu-

cose to glucose-6-phosphate for entry into glycolysis, carried significantly different

fluxes in three Arabidopsis models, irrespective of the biomass equation used. Based

on 10 flux units of glucose, the Poolman model produces 2.84 units of biomass by

using the biomass equation of Poolman model (i.e. PoolmanBOF), while the models

of AraGEM and AraCore synthesize 1.53 and 1.42 biomass units respectively. Sim-

ilarly, Poolman model yields the highest biomass by using AraGEMBOF and

AraCoreBOF (2.40 and 2.12 flux units, respectively), while, AraGEM generates the

lowest biomass (1.29 and 1.13 flux units, respectively; Figure 2.2). This indicates

that the Poolman model predicts more efficient conversion of glucose into biomass

compared to the other two models. Overall, model structure has a large impact on

the model prediction made with FBA, which highlights the importance of the quality

of metabolic models on the variation of model predictions.

To quantify the variation of fluxes in face of changing biomass composition

and model structure, we calculated the standard deviation, a robust measure for the

Sensitivity in plant metabolic models

29

dispersion within a set of data, for each reaction fluxes that were predicted from

nine model-biomass equation combinations. We then compared the median of the

standard deviation (Appendix 2.6.5). The medians of the standard deviations of re-

action fluxes for the cases with the same biomass composition but different models

(4.59, 3.35 and 3.22 for PoolmanBOF, AraGEMBOF and AraCoreBOF combina-

tions, respectively) were systematically higher than that of simulations with the

same metabolic network (0.29, 0.53 and 1.98 for Poolman, AraGEM and AraCore

combinations). This indicates that the predicted fluxes were considerably more dis-

persed when using different models, despite of the same biomass equation, revealing

that model structure has a larger impact on the FBA predictions.

For further confirmation, FVA was performed to examine the flux capacity for

each reaction under the 9 model-biomass combinations (Appendix 2.6.6). FVA

analysis showed that 10 of 33 considered reactions (pyruvate dehydrogenase, reac-

tion 13; citrate synthase, reaction 14; succinate thiokinase, reaction 18; complex II,

reaction 19; glucose 6-phosphate dehydrogenase, reaction 22; 6-phosphogluco-

nolactonase, reaction 23; ribulose-phosphate 3-epimerase, reaction 26; ribulose 5-

phosphate epimerase, reaction 27; transketolase 1, reaction 28; and transketolase 2,

reaction 29) have non-overlapping flux variability ranges in the compared scenarios

and agreed with the FBA analysis. However, the rest have a large range of possible

flux values in, at least one of considered scenarios, resulting in overlapping flux

variability ranges in the compared scenarios, which are not comparable.

Robustness of growth rate predictions with respect to changes in individual

biomass component and maintenance cost

From Figure 2.2, we observed that growth rate is not sensitive to the biomass com-

position. Considering the biomass objective function is the employed objective in

this study, which is the most commonly used objective for FBA (Feist and Palsson,

2010), we further analysed the impact of changing the fractional contribution of

each single biomass component on the growth rates with the ‘reference’ scenarios -

Poolman-PoolmanBOF, AraGEM-AraGEMBOF and AraCore-AraCoreBOF three

scenarios. The coefficient of each compound in the biomass equation was inde-

pendently varied 30% up or down (this value referred to Pinchuk et al., 2010) in

each ‘reference’ scenario, while the composition of the rest of the biomass compo-

nents was kept unchanged. The resulting model predictions of the growth rates were

compared with that of ‘reference’ biomass equation. Overall, our analyses showed

that the predicted growth rates were not sensitive to changes in the ratios of biomass

Chapter 2

30

components. For instance, growth rates in AraGEM and AraCore models were al-

tered by, at most, 4% and 6%, respectively (Appendix 2.6.7). The only exception is

in the Poolman model where a 30% decrease or increase of the composition of cell

wall led to 19.9% and 14.3% variation in growth rate, respectively (Appendix 2.6.7).

This discrepancy is not unexpected as cell wall is the largest part (66.6%) of the

overall biomass composition in the Poolman model (Figure 2.1).

Our analysis preferentially revealed that growth rate has a larger change at

higher coefficient in terms of C atom. Given that glucose, which is the carbon and

energy source, was used as the limiting nutrient in the analyses in this study, we

would expect growth rate correlates to the fractional coefficient of C atom. To quan-

tify this effect, we calculated the Pearson correlation coefficient (PCC), a measure

of the linear dependence between two variables, for all predicted growth rates sim-

ulated by decreasing and increasing of each single biomass component by 30%

(Appendix 2.6.8). We also calculated the correlation coefficients between changes

in growth rate and the coefficient in terms of weight (Appendix 2.6.9). We found

that the correlations among the coefficient in terms of weight were systematically

similar to the observed correlations among the coefficient in terms of C atoms (Ap-

pendix 2.6.8), indicating that when the fractional contribution of C atoms were

altered, the growth rates were increased or decreased in a synchronized fashion.

In addition to the structure of the metabolic network and biomass composition,

the in silico growth rate can also be influenced by the maintenance cost, i.e. GAM

and NGAM. Therefore, we evaluated the influence of maintenance on the predicted

growth rates with three ‘reference’ scenarios characterized by constraints on: (i) no

maintenance, (ii) sole GAM, (iii) sole NGAM, and (iv) GAM and NGAM. We then

compared the growth rates of scenario (i), (ii) and (iii) with that of scenario (iv),

separately. The analyses showed that the influence of maintenance parameters var-

ies with the used model. The exclusion of GAM and/or NGAM changes the growth

rate by, at most, 29.6% and 1.3% in AraGEM model and AraCore model, respec-

tively (Appendix 2.6.10). Given that AraGEM and AraCore models describe the

ETC reactions quite differently (section 2.3.1), it is unsurprising to observe large

discrepancies between the influences of maintenance cost on growth rate predictions

for these two models. It is worth mentioning that the growth rate predicted from the

Poolman model did not change with the maintenance, which is biologically implau-

sible. This is because there exist futile cycles in Poolman model (Arnold et al.,

2015).

Sensitivity in plant metabolic models

31

2.4 Discussion

In this chapter, we investigated the effect of biomass composition and model struc-

ture on steady-state flux predictions in the metabolic models of Arabidopsis.

Different combinations (3 models and 3 biomass equations) were tested to cover

multiple possibilities for biomass composition. Our results demonstrated that flux

predictions of the central metabolic network are fairly insensitive to changes in bi-

omass composition, regardless of the employed metabolic networks (Figure 2.2).

Biomass demands are a principal drive of the flux distributions in metabolic models,

therefore one would expect the flux distributions to be sensitive to changes in the

biomass composition. However, several experimental studies provided evidence

that central carbon metabolism is rather stable when the substrate source remains

unchanged (Rontein et al., 2002; Spielbauer et al., 2006; Junker et al., 2007; Willi-

ams et al., 2008; Williams et al., 2010). In a study of heterotrophic Arabidopsis cell

culture, fluxes through the central metabolic network measured by metabolic flux

analysis (MFA) were observed to be unchanged under different physiological con-

ditions, despite under which the biomass composition altered (Williams et al.,

2008). Consistently, Spielbauer et al. (2006) found that flux distributions of central

carbon metabolism were stable in maize endosperm. Nonetheless, under conditions

with different substrates provided, the flux distribution in the central carbon metab-

olism altered significantly (Junker et al., 2007). The noticeable stability of central

carbon metabolism has been extensively demonstrated in many microorganism

studies (Pramanik and Keasling, 1997; Pramanik and Keasling, 1998; Feist et al.,

2007).

Our results also showed that 30% variations in individual biomass precursor

had a minor effect on the growth rates with Arabidopsis models (Appendix 2.6.7).

This observation is in agreement with the previously reported findings (Pramanik

and Keasling, 1998; Puchałka et al., 2008). Puchalka et al. (2008) reported that var-

ying a single biomass component by 20% up or down has a negligible effect (less

than 1%) on the growth rate. Due to the large number of macromolecules such as

protein and carbohydrate in the cell, it is not surprising that single biomass compo-

nent, which makes up the macromolecules did not significantly affect the growth

rate. However, in other studies, it was found that varying the macromolecular com-

position had considerable effect on the flux simulations (Feist et al., 2007; Nookaew

et al., 2008). Similarly, our analyses showed that the effect of maintenance cost on

growth rate was limited (Appendix 2.6.10), which contradicts previous results of E.

coli model that maintenance cost had larger impact on growth rate prediction (Feist

Chapter 2

32

et al., 2007). However, it is important to point out that the growth yield altered

merely by 5% when increasing or decreasing GAM twofold in the study of Puchalka

et al. (2008), whereas, the effect of NGAM depends on the rate of carbon source

supply. The discrepancies of the effect of maintenance cost on growth rate predic-

tions of models of different organisms is likely to due to the difference in lifestyle

and metabolic behavior of the different organisms.

Sensitivity analyses of biomass composition performed on different models of-

ten yields conflicting results. According to the sensitivity analysis on the E. coli

model of Feist et al. (2007), small changes in biomass composition did not signifi-

cantly affect growth rates, while, cellular maintenance cost can considerably

influence growth rate prediction. However, the study of the E. coli model of Varma

and Palsson (1995) noted that predictions of constraint-based model are not very

sensitive to maintenance, including GAM and NGAM. Recently, Dikicioglu et al.

(2015) found that the predicted fluxes of the metabolic models of yeast were sensi-

tive to biomass composition. These are likely caused by inherent differences in the

respective metabolic networks. To verify this, we examined the differences of FBA

solutions between different models by constraining the same biomass outputs and

cellular maintenance costs, and our results indicated that model structure has a larger

impact on the flux predictions than the biomass composition (Figure 2.2). Moreover,

our analyses of the impact of maintenance on the predicted growth rates showed

large differences between the three examined models that represent the ETC reac-

tion differently (Appendix 2.6.10). These observations indicate the structure of

metabolic models is a major determinant of the variation in model predications.

Although our analysis indicated the robustness of the growth rate and central

metabolic fluxes in response to changes in biomass composition and maintenance

cost, it does not suggest that it is sufficient to generalize biomass composition from

any cell type within a plant or from other closely related organisms. This is because

the biomass composition for different tissues and organisms may include metabo-

lites specific to them, which need specific determination. Notably, variables other

than model structure, such as the P/O ratio have considerable impacts on flux dis-

tributions according to previous reports (Feist et al., 2007).

2.5 Conclusions

In this chapter, we reviewed the characteristics of published Arabidopsis metabolic

models, concentrated on three particular models with distinct model structures and

different compositions of biomass components. We examined the sensitivity of pre-

dicted fluxes to biomass composition, which is particularly relevant to plant

Sensitivity in plant metabolic models

33

metabolic models for which the biomass data is often collected from various

sources. Our analysis showed that central metabolic fluxes as well as growth rates

were insensitive to the variations in biomass composition, but were significantly

affected by model structure. This work represents a thorough set of analyses per-

formed in plants by means of constraint-based modelling, thereby providing

relevant information about how critical FBA solutions can be affected by biomass

composition, and more importantly by the structure of the models. Despite differ-

ences in several aspects such as model structure and number of metabolites and

reactions included, each of the evaluated models has its own merits. Comparative

analysis of the models paves the way for exploring the existence of principles that

are relevant for the regulation and robustness of plant central carbon metabolism.

2.6 Appendix

Chapter 2

34

Appendix 2.6.1 Newly added transporters in three Arabidopsis models

Model ReactionID Reaction Name Reaction Equation Classification

Poolman GLC_bm_tx' Glucose biomass transporter GLC -> Biomass transporter

Suc_bm_tx' Sucrose biomass transporter SUCROSE -> Biomass transporter

'Fru_bm_tx' Fructose biomass transporter FRU -> Biomass transporter

'Trehalose_bm_tx' Trehalose biomass transporter TREHALOSE -> Biomass transporter

'Succinate_bm_tx' Succinate biomass transporter SUC -> Biomass transporter

'Fumarate_bm_tx' Fumarate biomass transporter FUM -> Biomass transporter

'Malate_bm_tx' Malate biomass transporter MAL -> Biomass transporter

'Shikimate_bm_tx' Shikimate biomass transporter SHIKIMATE -> Biomass transporter

'Urea_bm_tx' Urea biomass transporter UREA -> Biomass transporter

H2S_tx' H2S transporter -> HS Nutrient uptake

AraGEM 'Ornithine_bm_tx' Ornithine biomass transporter S_L_45_Ornithine[c] -> S_Ornithine_biomass Biomass transporter

'Trehalose_bm_tx' Trehalose biomass transporter S_alpha_44_alpha_45_Trehalose[c] -> S_Trehalose_biomass Biomass transporter

'Succinate_bm_tx' Succinate biomass transporter S_Succinate[c] -> S_Succinate_biomass Biomass transporter

'Fumarate_bm_tx' Fumarate biomass transporter S_Fumarate[c] -> S_Fumarate_biomass Biomass transporter

'Malate_bm_tx' Malate biomass transporter S__40_S_41__45_Malate[c] -> S_Malate_biomass Biomass transporter

'Shikimate_bm_tx' Shikimate biomass transporter S_Shikimate[c] -> S_Shikimate_biomass Biomass transporter

'Urea_bm_tx' Urea biomass transporter S_Urea[c] -> S_Urea_biomass Biomass transporter

AraCore 'Im_Glc' Glucose exchange -> Glc[c] Nutrient uptake

Sensitivity in plant metabolic models

35

Appendix 2.6.2 The calculation of the weight percentage of each biomass

constitute. Can be found online at: http://journal.frontiersin.org/arti-

cle/10.3389/fpls.2016. 00537

Appendix 2.6.3 Schematic representation of the central carbon metabolic

network and related metabolism.

Each reaction is numbered, referencing Appendix 2.6.4. Metabolite abbreviations are as follows: GLC, glucose; 2-

OG, 2-oxoglutarate; 3-PGA, 3-phosphoglycerate; 2-PG, 2-phosphoglycolate; G6P, glucose-6-phosphate; F6P, fruc-

tose-6-phosphate; 6PGL, 6-phosphogluconolactone; 6PG, 6-phosphogluconate; Ru5P, ribulose-5-phosphate; R5P,

ribose-5-phosphate; X5P, xylulose-5-phosphate; S7P, sedoheptulose-7-phosphate; E4P, erythrose-4-phosphate;

FBP, fructose-1,6-biphosphate; DHAP, dihydroxyacetone phosphate; DPG, glycerate-1,3-bisphosphate; PEP,

phosphoenol pyruvate; OAA, oxalacetic acid; GAP, glyceraldehyde-3-phosphate; 2-PGA, 2-phosphoglycerate;

Pyr, pyruvate; Cit, citrate; IsoCit, threo-isocitrate; Suc, succinate; SucCoA, succinyl-CoA; Fum, fumarate; Mal,

malate; QH2, ubiquinone; Q, ubiquinol; Cytred, cytochrome reduced; Cytox, cytochrome oxidized; Fdred, ferredoxin

reduced; Fdox, ferredoxin oxidized; BOF, biomass objective function; Ala, alanine; Arg, arginine; Asn, asparagine;

Asp, aspartate; Cys, cysteine; Gln, glutamine; Glu, glutamate; Gly, glycine; Ile, isoleucine; Leu, leucine; Lys, ly-

sine; Met, methionine; Phe, phenylalanine; Pro, proline; Ser, serine; Thr, threnine; Tyr, tyrosine; Val, valine.

Chapter 2

36

Appendix 2.6.4 Central carbon metabolic reactions included in three Ara-

bidopsis models. Can be found online at:

http://journal.frontiersin.org/article/10.3389/fpls.2016. 00537

Appendix 2.6.5 The calculation of the standard deviation for scenarios. Can

be found online at: http://journal.frontiersin.org/article/10.3389/fpls.2016.

00537

Appendix 2.6.6 FVA results. Can be found online at: http://journal.fron-

tiersin.org/article/10.3389/fpls.2016. 00537

Appendix 2.6.7 Influence of individual biomass component on the growth

rate. Can be found online at: http://journal.frontiersin.org/arti-

cle/10.3389/fpls.2016. 00537

Appendix 2.6.8 The correlation analysis between the change of growth rate

and the coefficient in terms of C atoms.

(a) Poolman, (b) AraGEM and (c) AraCore model. The Pearson Pearson correlation coefficient was calculated

using the changes in growth rate for + 30% variations of a single biomass component.

Sensitivity in plant metabolic models

37

Appendix 2.6.9 The correlation analysis between the change of growth rate

and the coefficient in terms of weight.

(a) Poolman, (b) AraGEM and (c) AraCore model. The Pearson Pearson correlation coefficient was calculated

using the changes in growth rate for + 30% variations of a single biomass component.

Appendix 2.6.10 Influence of maintenance on the growth rate.

Scenarioes % Change in growth rate

No maintenance GAM NGAM

Poolman-PoolmanBOF 0 0 0

AraGEM-AraGEMBOF 29.59 0.66 28.75

AraCore-AraCoreBOF 1.34 0.57 1.34

Chapter 3 Constructing a genome-

scale metabolic model for tomato1

1 Part of this this chapter has been published:

Yuan, H., Cheung, C.Y.M., Poolman, M.G., Hilbers, P.A.J. and van Riel,

N.A.W. (2016) A genome-scale metabolic network reconstruction of to-

mato (Solanum lycopersicum L.) and its application to photorespiratory

metabolism. Plant J., 85, 289–304.

Abstract

Tomato (Solanum lycopersicum L.) has been studied extensively owing to its high

economic value in the market, and high content in health-promoting antioxidant

compounds. Tomato is also considered as an excellent model organism for studying

the development and metabolism of fleshy fruits. In this chapter, we reconstructed

iHY3410, a genome-scale metabolic model of tomato, which represents a growing

leaf cell. The resulting model consists of 3,410 genes and 2,143 biochemical and

transport reactions distributed across five intracellular organelles including cytosol,

plastid, mitochondrion, peroxisome and vacuole. The model can be used to simulate

cellular metabolic behaviors under heterotrophic and phototrophic conditions.

Constructing a genome-scale metabolic model for tomato

41

3.1 Introduction

Tomato (Solanum lycopersicum L.) is one of the most important vegetable crops

grown in the world. An estimated worldwide production is 161.8 million tons fresh

weight on 4.8 million hectares each year (FAO, 2014). In addition to its high eco-

nomic value in the market, tomato has always drawn much attention owing to its

high content in health-promoting antioxidant compounds such as ascorbic acid (vit-

amin C), carotenoids, and flavonoid (Dixon, 2005; Story et al., 2010). Among the

plentiful plant species bearing fleshy fruits, tomato serves as an excellent model

system for addressing the fruit development and metabolism because of its ease of

cultivation, and short life cycles (Klee and Giovannoni, 2011). Of particular interest

is that the tomato fruit undergoes a shift from partially photosynthetic to truly het-

erotrophic metabolism during development (Kahlau and Bock, 2008; Lytovchenko

et al., 2011). Furthermore, tomato plant has a number of distinct developmental fea-

tures, such as compound leaves and the branching behavior of its inflorescence,

which make it an attractive system to study desirable traits. Therefore, tomato is

considered to be an ideal model plant for both basic and applied studies.

In the past two decades, much effort has been made towards understanding the

metabolic properties of tomato, ranging from physiology and biochemistry to mo-

lecular and genetics by conventional experimental techniques (Haupt-Herting et al.,

2001; Lieberman et al., 2004; Liu et al., 2004). However, its metabolic behavior

and the interactions between different metabolic pathways across the whole system

are still not completely understood. A recent emerging modelling approach, namely

constraint-based modelling, has shown to be particularly promising as it offers a

powerful tool to systematize our current knowledge of the complex metabolic net-

work (Bordbar et al., 2014). Flux balance analysis (FBA), a constraint-based

modelling approach, is able to predict the metabolic fluxes with stoichiometric met-

abolic models, some of which are at genome-scale (Orth et al., 2010). Genome-scale

metabolic models (GSMs) are constructed by extracting the metabolic reactions en-

coded by the genomes of the target organisms. Large-scale metabolic models have

been reconstructed for a variety of plants thus far (Figure 3.1), including Arabidop-

sis (Poolman et al., 2009; de Oliveira Dal’Molin et al., 2010a; Cheung et al., 2013;

Arnold and Nikoloski, 2014), barley (Grafahrend-Belau et al., 2009; Grafahrend-

Belau et al., 2013), rice (Poolman et al., 2013), oilseed rape (Hay and Schwender,

2011), and C4 plants (de Oliveira Dal’Molin et al., 2010b; Saha et al., 2011; Simons

et al., 2014). Most recently, a stoichiometric model was developed to describe the

Chapter 3

42

developing process of tomato fruit by modelling central metabolism (Colombié et

al., 2015), but at a medium-scale level. Nevertheless, no large-scale model of tomato

metabolism has been published to date. With the recent completion of the tomato

genome in 2012 (Tomato Genome Consortium, 2012), it is timely to expand the

scope to a genome-scale model of tomato metabolism for better understanding its

metabolic behavior by integrating biochemical, physiological, and proteomic data

derived from the literature and databases.

A general overview of the genome-scale metabolic reconstruction has been

given in Chapter 1 (section 1.3.1). In this chapter, we describe the reconstruction of

a compartmentalized genome-scale metabolic model for tomato, we then use this

model to simulate tomato leaf metabolism. We evaluated the model by simulating

several well-studied metabolic scenarios and showed that the simulation results

were in good agreement with the main biochemical properties of plant metabolism

as described in the literature.

3.2 Methods

3.2.1 Software

In this thesis work, all modeling work of tomato was carried out using ScrumPy,

which is an open source, metabolic modeling software package written in Python

programming language (Poolman, 2006). For detailed instructions of ScrumPy and

its usage, it is available from the main ScrumPy website

(http://mudshark.brookes.ac.uk/index.php/Software/ScrumPy).

3.2.2 Metabolic reconstruction

In this chapter, an initial draft model of tomato was generated from a tomato meta-

bolic pathway database LycoCyc (version 3.0;

http://pathway.iplantcollaborative.org/lycocyc.html) released by the Solanaceae

Genomics Network (SGN, http://sgn.cornell.edu/). This database is available as part

of BioCyc collection of databases (Karp et al., 2005). The draft model was manually

curated with regard to atomic balance, energy conservation, reaction stoichiometry,

reversibility, and subsequently assigned to respective subcellular localization. Mass

conservation was achieved by checking the atomic balance of each reaction for

which the empirical formulas of all metabolites are known, and by identifying un-

conserved metabolites through stoichiometric consistency analysis (Gevorgyan et

Constructing a genome-scale metabolic model for tomato

43

al., 2008). All reactions were atomically balanced with respect to carbon (C), nitro-

gen (N), phosphorus (P), and sulfur (S), but not hydrogen (H), and oxygen (O). The

inconsistency regarding H and O was assumed to have minimal effect on the model

predictions as water and stoichiometric protons were defined as external metabolites

as in Poolman et al. (2009) and Cheung et al. (2013).

Information on reaction directionality was primarily extracted from the Meta-

Cyc database (http://metacyc.org) wherever possible as it contains comprehensive

metabolic pathways and enzymes which have been experimentally determined and

reported in the literature (Caspi et al., 2008). For the remaining reactions, the direc-

tionality data was adopted from the LycoCyc database. Intracellular

compartmentalization was determined according to a published genome-scale

model of Arabidopsis (Cheung et al., 2013). Reactions were distributed in 5 intra-

cellular compartments, cytosol (‘_Cyto’), plastid (‘_Plas’), mitochondrion

(‘_Mito’), peroxisome (‘_Pero’) and vacuole (‘_Vacu’). Transport reactions were

also defined based on the model of Cheung et al. (2013). Transporters were repre-

sented using the subscripts ‘_Cyto_tx’, ‘_Plas_tx’, ‘_Mito_tx’, ‘Pero_tx’ and

‘_Vacu_tx’, for exchange with the environment (also termed exchange reactions

elsewhere), plastidic, mitochondrial, peroxisomal and vacuolar transporters, respec-

tively. To enable biomass production from the given substrates, biomass drains were

assigned using the subscript ‘_bm_tx’ independently accounting for protein, sugars,

nucleotides, soluble metabolites, cell wall, fatty acids, and pigments (Appendix

3.5.1).

The inclusion of protons in the model was defined as ‘Pumped-PROTON’ and

‘PROTON’, which represent energetic protons and stoichiometric protons, respec-

tively, in accordance with the published Arabidopsis models (Poolman et al., 2009;

Cheung et al., 2013). Pumped-PROTON is used to drive ATP synthesis and/or me-

tabolite transfer across membranes; PROTON is the stoichiometric protons, which

is regarded as external in all compartments.

3.2.3 Model analysis

In this chapter, flux balance analysis (FBA) was carried out using the Gnu linear

programming kit (GLPK, http://www.gnu.org/software/glpk) and the ScrumPy met-

abolic modelling package (Poolman, 2006). The objective function of total flux

minimization was used in this study as a proxy for minimizing the total enzyme

costs (Holzhütter, 2004), subject to several different types of constraints: (i) steady

Chapter 3

44

state, (ii) the requirement for production of individual biomass components (Appen-

dix 3.5.1), and (iii) the energy demand for cell maintenance. Mathematically, the

linear programming problem can be represented as follows:

Minimize:|v|

Subject to Sv = 0 (i)

vi = t (ii)

vATPase = 𝑚 (iii)

where S is the stoichiometric matrix, v is a vector of all reaction fluxes in the system,

i are the transporters of the biomass components, t is the corresponding flux that

each biomass transporter carries, vATPase represents the flux that a generic ATPase

reaction carries to sustain cellular functions that are not associated with biomass

accumulation and growth, and 𝑚 is the flux value that the generic ATPase reaction

carries. It has been shown that a large amount of cellular ATP is used for cellular

maintenance in plant cells, for example, more than 85% of cellular ATP was esti-

mated for the non-growth associated maintenance in Arabidopsis (Masakapalli et

al., 2010). Since no information is available regarding the ATP maintenance energy

for tomato, the ATP maintenance of 7.1 mmol g-1 DW h-1 reported in an Arabidopsis

model has been exploited (Poolman et al., 2009). As most plants uptake nitrogen in

the form of nitrate and ammonium, nitrogen input was set to be 50% nitrate and

50% ammonium for simulating the photoautotrophic scenario. Photorespiratory me-

tabolism was simulated by additionally setting the flux ratio of carboxygenic to

oxygenic of RubisCO (Vc/Vo), which is 3 under ambient conditions. Other con-

straints for simulating the photoautotrophic scenario are listed in Appendix 3.2.

FBA uses these constraints to identify a flux distribution which optimizes a defined

objective function (Orth et al., 2010), such as total flux minimization.

FBA was also applied to test the production of each biomass component inde-

pendently, while minimizing total flux to simulate growth. To characterize the

optimal flux solution space, flux variability analysis (FVA) was used to determine

a possible flux range for each reaction while giving rise to the same optimal value

for the objective function (Mahadevan and Schilling, 2003). This range denotes the

minimum and maximum flux values each reaction can take among all alternative

optimal solutions.

Constructing a genome-scale metabolic model for tomato

45

3.3 Results and Discussion

3.3.1 General model properties

In this chapter, we aimed to build a genome-scale metabolic network for tomato by

making use of the recent availability of the tomato genome (Tomato Genome Con-

sortium, 2012). The metabolic reconstruction was based on information in LycoCyc

(http://pathway.iplantcollaborative.org/lycocyc.html), published literature and

known biochemical knowledge in plants from textbook. The resulting model,

iHY3410, comprises 3,410 genes, 1,998 metabolites and 2,143 reactions, of which

1,211 could carry a non-zero flux given the defined nutrients input and the specified

biomass components. Among the 2,143 reactions, 1,885 represent biochemical con-

versions and 258 represent metabolite transporter processes indicating the high

interconnectivity among the intracellular compartments (i.e. cytosol, mitochondria,

plastid, peroxisome, and vacuole) and with the environment. There were 11 ex-

change reactions with the environment, of which O2 and CO2 can freely exchange

with the environment. Biomass synthesis was represented by a set of independent

transporters, one for each representative of biomass components: protein (amino

acids), sugars, nucleotides, soluble metabolites, cell wall (cellulose, xylan and lig-

nin), fatty acids and pigments (chlorophyll a and chlorophyll b). The model was able

to simulate biomass synthesis of leaf using starch or CO2, NO3- and/or NH4

+, SO42-,

and Pi as the C, N, S, and P sources, respectively. A graphical overview of the re-

constructed network is shown in Figure 3.1.

Figure 3.1 Schematic overview of iHY3410. Different compartments are indicated along with the number of asso-

ciated reactions. Grey arrows denote transporters across membranes or as biomass drains, with the number of

reactions indicated.

Chapter 3

46

The definition of the photosynthetic light reactions in iHY3410 was consider-

ably expanded over previous models of other plant species (de Oliveira Dal’Molin

et al., 2010a; Lakshmanan et al., 2013; Poolman et al., 2013). Rather than using

aggregate reactions for the photosynthetic electron transport, we provided sufficient

detail in iHY3410 to define all individual electron transport reactions including pho-

tosystem II (PSII), cytochrome b6f complex, photosystem I (PSI), and ferredoxin

NADP+ reductase (FNR) as components of the linear electron chain, the plastidic

ATP synthase and the cyclic electron transfer ferredoxin plastoquinone reductase

(FQR) reaction around PSI. The reconstruction was provided in SBML (System Bi-

ology Markup Language (Hucka et al., 2003)) format, ScrumPy format and Excel

format Appendix 3.5.3, 3.5.4 and 3.5.5, respectively. The inclusion of gene-protein-

reaction (GPR) associations enables iHY3410 to be used for the integration and

contextualization of proteomic and transcriptomic data.

3.3.2 Comparisons with existing plant models

The model properties of iHY3410 were compared with genome-scale metabolic

models of two plant species, Arabidopsis (Cheung et al., 2013) and rice (Poolman

et al., 2013). Our tomato model contained more unique reactions and metabolites

than the rice model but fewer than the Arabidopsis model. One possible reason for

the higher number of reactions and metabolites in the Arabidopsis model is that the

biochemical databases from which the construction of the Arabidopsis model is

based on is more comprehensively annotated. Of all the unique reactions and me-

tabolites in iHY3410, 1,417 reactions (72% overall) and 1,344 metabolites (81%

overall) were also present in Arabidopsis model (Figure 3.2a). Similarly, we sys-

tematically compared iHY3410 to a rice model (Figure 3.2b). Our tomato model

shared 1,155 reactions (64% overall) and 1,043 metabolites (63% overall) with the

rice model. This overlap was smaller than that with the Arabidopsis model, which

reflects the divergence between eudicots, including Arabidopsis and tomato, and

monocots such as rice (Considine et al., 2002; Tomato Genome Consortium, 2012).

Irrespective of the Arabidopsis or rice models, a total of 269 reactions and 240 me-

tabolites are specific to our tomato model. These reactions were predominately

involved in secondary metabolism such as volatiles, alkaloids and phytoalexins bi-

osynthesis. For example in iHY3410, reactions 'RXN-6721', '2.1.1.146-RXN',

'RXN-6741' and 'RXN-6742' described the biosynthesis of volatile ester, which has

no counterpart in Arabidopsis or rice. While iHY3410 was applied to model a grow-

ing tomato leaf cell in this chapter, the inclusion of reactions that are dedicated to

Constructing a genome-scale metabolic model for tomato

47

the tomato fruits enables future investigations and characterization of fruit metabo-

lism using our model. Phytoalexins have been considered as plant antibiotics in the

plant defense systems, which are induced and accumulated in plants in response to

microbial infection or abiotic stress (Kuc, 1995). Our tomato model contains a se-

quence of four reactions, 'RXN-4843’, 'RXN-4823','RXN-4844' and 'RXN-4848',

which are responsible for sesquiterpenoid phytoalexins biosynthesis. Although phy-

toalexins are not synthesized in healthy plants (Back et al., 1998), the inclusion of

these reactions provides a platform for the future in silico investigation of disease

states in tomato.

Figure 3.2 Venn diagram for reactions and metabolites of different genome-scale metabolic models for plants. (a)

The number of unique reactions and metabolites shared between and specific for the Arabidopsis model of Cheung

et al. (2013) and iHY3410, (b) The number of unique reactions and metabolites shared between and specific for the

rice model of Poolman et al. (2013) and iHY3410. All of the models were de-compartmentalized so that only the

unique reactions and the cytosolic transporters including biomass transporters were compared. The numbers in the

black shaded areas indicate the number of metabolites or reactions unique to iHY3410, while the numbers in the

empty areas represent the number of metabolites or reactions unique to the Arabidopsis or rice models, the numbers

in the grey regions represent the number of common metabolites or reactions between iHY3410 and Arabidopsis

or rice.

3.3.3 Model validation

The functional capabilities of iHY3410 were examined by predicting cell growth

under heterotrophic, phototrophic and photorespiratory conditions using FBA with

minimization of total flux as the objective function subjecting to mass-balance con-

straints for all internal metabolites and to the flux constraints of biomass production

and ATP maintenance cost (see 3.2.2 Metabolic reconstruction). In total, between

Chapter 3

48

359 and 371 (16.8%-17.3%) of all reactions carried a non-zero flux in the examined

conditions, while the majority of the reactions were not used in the simulations. Our

observation did not stand alone as similar observations had also been reported in

previous studies. For example, only 232 of 1,406 reactions (16.5%), 309 of 1,736

reactions (17.8%), and 248 out of 1,097 reactions (22.6%) were used for biomass

synthesis in a genome-scale model of Arabidopsis (Poolman et al., 2009), rice

(Poolman et al., 2013), and Salmonella. Typhimurium (Hartman et al., 2014), re-

spectively. Given that secondary metabolites were not considered in the biomass

constraints, it is not surprising that only a subset of reactions were active in our

model prediction.were not considered in the biomass constraints, it is not surprising

that only a subset of reactions were active in our model prediction.

Heterotrophic conditions

During the dark period, cellular metabolism of leaves is supported by the catabolism

of stored compounds that are accumulated during the light period. In this study, the

metabolic behavior of a growing tomato leaf was simulated by considering starch

as the sole carbon source in the dark. As expected, the model predicted that starch

was degraded to form maltose and glucose-1-phosphate (G1P). The glucose-6-phos-

phate (G6P) generated from G1P was predicted to be oxidized via the oxidative

pentose phosphate (OPP) pathway, which predominately provides reducing power

for the production of biomass, in particular for fatty acid synthesis (Figure 3.3a).

The model also predicted that the mitochondrial pyruvate dehydrogenase (PDH) is

active in darkened leaves, providing acetyl-CoA for fatty acid biosynthesis and as a

substrate for the TCA cycle. Thereby, the TCA reactions operate in a conventional

cyclic manner. The TCA cycle functions not only to generate ATP and reducing

equivalents but also to provide carbon skeletons for the biosynthesis of amino acids.

The operation of a complete cyclic flux through the TCA cycle was thought to be

favorable to provide efficient energy in the dark, which is in line with the previous

reports (Tcherkez et al., 2005).

Phototrophic conditions

In photoautotrophic growth conditions, biomass synthesis is supported by the as-

similation of CO2 by the enzyme ribulose-1, 5-bisphosphate carboxylase/oxygenase

(RubisCO) using light as the energy source. Under this imposed mass balance and

with the fixed biomass, minimization of total flux predicted a photon influx of 9.84

mmol g-1 DW h-1. The production of ATP and reducing power by the photosynthetic

Constructing a genome-scale metabolic model for tomato

49

light reaction liberates OPP pathway from supplying reducing power. The flux map

obtained under illumination therefore exhibited a completely different flux distribu-

tion compared with heterotrophic conditions (Figure 3.3b). Not surprisingly, the

pentose phosphate pathway shifted to the reductive mode as ribulose-5-phosphate

kinase (PRK) and RubisCO became active in the presence of light, which then drove

the subsequent reactions of the Calvin cycle.

Compared with the cyclic mode of the TCA cycle in heterotrophic conditions,

the TCA cycle reactions shifted to a noncyclic mode in phototrophic conditions in

our prediction, where the reactions from α-ketoglutarate (2-OG) to malate carry no

flux. The enzyme 2-oxoglutarate dehydrogenase (OGDH) plays an essential role in

overall metabolic activity, in particular nitrogen assimilation (Bunik and Fernie,

2009). Nevertheless, the genetic manipulation of OGDH in tomato plants revealed

little impact on the photosynthetic capacity (Araújo et al., 2012). Given the fact that

multiple cytosolic isoforms exist for the mitochondrial TCA reactions, it would not

be surprising that some mitochondrial isoforms were not used in our model predic-

tion (Sweetlove et al., 2010), e.g. fumarase operates in the form of cytosolic bypass

in the prediction. Succinate dehydrogenase (SDH, also known as Complex II) has a

dual function as a component of the TCA cycle and the electron transport chain

(ETC). As yet, the role of SDH has been studied in tomato plants employing muta-

genic approach, showing that inhibition of the iron-sulfur subunit of Complex II had

a favorable effect on the photosynthesis (Araújo et al., 2011). Thus, it is not unex-

pected that in our prediction, Complex II was not used under light conditions (Figure

3.3b). The operation of an incomplete TCA cycle in the light by our tomato model

is supported by a study in French bean leaves (Phaseolus vulgaris), which indicated

that the TCA cycle was almost completely inhibited in the light (Tcherkez et al.,

2005). Similarly, Poolman et al. (2013) reported that the TCA cycle did not function

as a conventional cycle at high light conditions in rice.

Photorespiratory conditions

To investigate specific features of photorespiratory metabolism in tomato plants, we

simulated the reconstructed model by setting a 3:1 carboxylation to oxygenation

ratio (Vc/Vo=3) based on experimental evidence (Leegood, 2007), with other con-

straints including biomass synthesis unchanged. Not surprisingly, the photon input

was 10.89 mmol g-1 DW h-1, higher than that of non-photorespiratory conditions

(9.84 mmol g-1 DW h-1) as photorespiration is generally thought of as a process

Chapter 3

50

Figure 3.3 Flux distribution of the tomato central metabolism under (a) dark conditions and (b) light conditions.

The solid lines indicate reactions with non-zero fluxes, the grey lines indicate reactions with zero fluxes, and the

dashed lines indicate reactions with zero or non-zero fluxes based on FVA solutions.

(a)

(b)

Constructing a genome-scale metabolic model for tomato

51

which reduces the efficiency of carbon fixation (Douce and Neuburger, 1999). The

resulting flux distribution is depicted in Figure 3.4.

Figure 3.4 Flux distribution of photorespiratory metabolism under normal conditions (Vc/Vo=3).

The photorespiratory pathway spanned four subcellular compartments in

iHY3410 including the plastid, the peroxisome, the mitochondrion and the cytosol.

The associated transporters between different compartments such as glycolate and

glycerate transporters were also included in the model. It has been established that

the mitochondrial enzyme glycine decarboxylase (GDC), which catalyzes the tetra-

hydrofolate-dependent catabolism of glycine, has essential roles in metabolic

processes. This enzyme functions in the photorespiratory pathway of all photosyn-

thetic tissues of C3 plants. In addition, it contributes to one-carbon metabolism in

association with serine hydroxymethyltransferase (SHMT), which forms an obliga-

tory route for C1 metabolism (Mouillon et al., 1999; Hanson and Roje, 2001; Engel

et al., 2007). In our model prediction, the cytosolic SHMT was in operation with

GDC (reaction 7, Figure 3.4).

During photorespiration, ammonia is generated in the mitochondria due to the

oxidation of glycine by GDC. This is re-assimilated by the glutamine synthe-

tase/glutamate synthase (GS/GOGAT) system to scavenge the cytotoxic metabolite.

The mechanism of NH3 re-assimilation has been a subject of controversy to date.

Chapter 3

52

GOGAT, either in the form of Fd-GOGAT or NADH-GOGAT, is exclusively pre-

sent in the plastid to produce glutamate (Coschigano et al., 1998). Therefore, it was

assumed that the photorespiratory NH3 is transported to plastids, which is then re-

assimilated into glutamate by GS and GOGAT (Wingler et al., 2000). On the con-

trary, an isoform of GS was found to be dual targeted to the chloroplast and the

mitochondria in Arabidopsis leaves (Taira et al., 2004). Based on this, Linka and

Weber (2005) hypothesized a possible metabolite route where the NH3 is assimi-

lated by the mitochondrial GS, which does not require more energy than the plastidic

GS route. Our model prediction supports the hypothesis that NH3 generated by GDC

during photorespiration could be re-assimilated by the mitochondrial GS, rather than

being transported to the chloroplast.

For further confirmation, FVA was performed to examine the flux capacity for

each reaction under the simulated conditions (Appendix 3.5.6). As expected, light-

dependent reactions such as photosystem II and photosystem I in the photosynthetic

light reactions, PRK and RubisCO from the Calvin cycle, nitrate reductase (NR),

nitrite reductase (NiR), plastidic GS and Fd-GOGAT involved in nitrogen assimila-

tion were activated by light. These reactions characterize the metabolic response

that must be appropriately regulated in order to adapt to the available energy/carbon

source.

3.4 Conclusions

The aim of this chapter was to construct a genome-scale metabolic network for

tomato, representing a growing leaf cell. The GSM of tomato reconstructed in this

chapter was initially extracted from the LycoCyc database, then was manually

curated in an iterative cycle. In addition, the ability of the model was tested by

predicting cellular behaviours under hetertrophic, phototrophic and

photorespiratory growth conditions. Eventually, the resulting model, iHY3410, in-

cludes 3,410 genes and 2,143 biochemical and transport reactions distributed across

five intracellular organelles including cytosol, plastid, mitochondrion, peroxisome

and vacuole. The model successfully described the known metabolic behavior of

tomato leaf under heterotrophic and phototrophic conditions, which provides a

sound framework for investigating tomato metabolism.

3.5 Appendix

Appendix 3.5.1 Detailed information on the biomass composition of to-

mato leaves.

Constructing a genome-scale metabolic model for tomato

53

Table 1: Concentration of biomass components in tomato leaves.

Component Concentration

(µmol/g FW)

Concentration

(mmol/g DW)

Reference

Amino acids

Alanine 1.26 1.48 × 10-2 Schauer et al., 2005

Arginine 0.24 2.82 × 10-3 Schauer et al., 2005

Aspargine 0.23 2.71 × 10-3 Schauer et al., 2005

Aspartate 6.36 7.48 × 10-2 Schauer et al., 2005

β-Alanine 0.11 1.29 × 10-3 Schauer et al., 2005

GABA 1.45 1.71 × 10-2 Schauer et al., 2005

Glutamate 11.51 0.135 Schauer et al., 2005

Glutamine 3.39 3.99 × 10-2 Schauer et al., 2005

Glycine 1.89 2.22 × 10-2 Schauer et al., 2005

Isoleucine 0.93 1.09 × 10-2 Schauer et al., 2005

Leucine 0.53 6.24 × 10-3 Schauer et al., 2005

Lysine 0.06 0.71 × 10-3 Schauer et al., 2005

Methionine 0.05 0.59 × 10-3 Schauer et al., 2005

Ornithine 0.31 3.65 × 10-3 Schauer et al., 2005

Phenylalanine 0.74 8.71 × 10-3 Schauer et al., 2005

Proline 18.59 0.219 Schauer et al., 2005

Serine 4.88 5.74 × 10-2 Schauer et al., 2005

Threonine 14.97 0.176 Schauer et al., 2005

Tryptophan 2.80 3.29 × 10-2 Schauer et al., 2005

Tyrosine 0.15 1.76 × 10-3 Schauer et al., 2005

Valine 1.10 1.29 × 10-2 Schauer et al., 2005

Sugars

Starch 42.73 0.503 Roessner-Tunali et al., 2003

Glucose 59.97 0.706 Schauer et al., 2005

Fructose 26.20 0.308 Schauer et al., 2005

Sucrose 16.70 0.196 Schauer et al., 2005

Nucletides

ATP 117.20 1.38 Roessner-Tunali et al., 2003

UTP 51.24 0.603 Roessner-Tunali et al., 2003

Soluble metabolites

Glucose-6-P 1.00 1.18 × 10-2 Schauer et al., 2005

Citrate 4.06 4.78 × 10-2 Schauer et al., 2005

Fumarate 0.14 1.65 × 10-3 Schauer et al., 2005

Glycolate 0.07 0.82 × 10-3 Schauer et al., 2005

Isocitarte 0.06 0.71 × 10-3 Schauer et al., 2005

α-Ketoglutarate 0.30 3.53 × 10-3 Schauer et al., 2005

L-Ascorbate 1.50 1.77 × 10-2 Schauer et al., 2005

Malate 28.93 0.340 Schauer et al., 2005

Succinate 2.16 2.54 × 10-2 Schauer et al., 2005

Glycerol 0.32 3.76 × 10-3 Schauer et al., 2005

Uracil 1.00 1.18 × 10-2 Schauer et al., 2005

5-Oxoproline 6.33 7.45 × 10-2 Schauer et al., 2005

Cell wall

Cellulose - 1.77 Sheen, 1983; Table 2

Hemicellulose (Xylan) - 1.74 × 10-2 Schauer et al., 2005; Table 2

Lignin

Coumaryl alcohol - 0.12 Sheen, 1983; Table 2

Coniferyl alcohol - 0.10 Sheen, 1983; Table 2

Chapter 3

54

Sinapyl alcohol - 0.0890 Sheen, 1983; Table 2

Fatty acids

FA 1.00 1.78 × 10-2 Schauer et al., 2005

Pigments

Chlorophyll a 4.1 4.82 × 10-2 Nunes-Nesi et al., 2007

Chlorophyll b 1.3 1.53 × 10-2 Nunes-Nesi et al., 2007

Note that the biomass compositions used here are experimentally determined from plant fresh weight, subsequently,

converted into dry weight (DW) based on the dry matter content (g DW/g FW, 0.085) of tomato (El-Sayed, 2013).

Since no information is available regarding the cellulose and lignin composition for tomato, the data were collected

from tobacco reported by Sheen (1983).

Table 2: Cellulose and lignin precursors.

Original data

(mg/g DW)

Average

(mg/g DW) WT %

MW

(g/mol) mmol/g DW

Cellulose (Glc dimer)

309.00

319.20 100 180.156 1.77

333.60

316.30

279.10

331.20

320.20

321.20

300.30

345.90

335.20

Lignin

55.90

56.10

40.70

81.00

Coumaryl alcohol 47.50 33.33 150.0681 0.12

Coniferyl alcohol 49.90 33.33 180.0786 0.10

Sinapyl alcohol 52.80 33.33 210.0892 0.0890

56.70

44.60

61.4

70.50

Table 3: Hemicellulose precursors.

Hemicellulose Components

Arabinose Xylose Mannose Galactose Galacturonate

µmol/ g FW 0.11 0.28 0.16 0.17 0.76

mmol/g DW 1.29 × 10-3 3.29 × 10-3 1.88 × 10-3 2.00 × 10-3 8.94 × 10-3

Total 1.74 × 10-2 mmol/g DW

Constructing a genome-scale metabolic model for tomato

55

Table 4: Condition-specific growth rates of tomato leaves.

Biomass concentrations (Table 3) are scaled to flux units (µmol/g DW/h) for normal and drought conditions, re-

spectively, based on the growth rates of tomato for normal and drought conditions (Sánchez-Rodríguez et al., 2010).

Other biomass precursors which are not listed in Table 3 were set to carry zero flux. ‘Original data’ refers to the

data derived from Sánchez-Rodríguez et al. (2010).

Appendix 3.5.2 Boundary constraints for simulating the photoautotrophic

scenario.

Columns marked with lb and ub represent the default lower and upper boundary, respectively.

NO3_Cyto_tx and NH4_Cyto_tx were constrained to a ratio of 1:1. ‘m’ is the flux value that the generic ATPase

reaction carries to sustain cellular functions that are not associated with biomass accumulation and growth.

Appendix 3.5.3 Genome-scale metabolic model of iHY3410 in SBML for-

mat. Can be found at:

http://onlinelibrary.wiley.com/doi/10.1111/tpj.13075/abstract

Items Original data

(mg/g DW/day)

Average

(mg/g DW/day) (mg/g DW/h)

Cultivar Kosaco Josefina Katalina Salomé Zarina

Normal 87.23 82.05 87.25 100.09 93.85 90.09 3.75

Drought 64.74 54.84 64.38 76.25 80.28 68.10 2.84

lb ub

Photon_tx 0 +infinity

NO3_Cyto_tx 0 +infinity

NH4_Cyto_tx 0 +infinity

SO4_Cyto_tx 0 +infinity

Pi_Cyto_tx 0 +infinity

K_Cyto_tx 0 +infinity

Ca_Cyto_tx 0 +infinity

Mg_Cyto_tx 0 +infinity

Fe_Cyto_tx 0 +infinity

O2_Cyto_tx -infinity +infinity

CO2_Cyto_tx -infinity +infinity

ATPase_Cyto m m

NADPHox_Cyto 0 +infinity

NADHox_Cyto 0 +infinity

NADPHox_Mito 0 +infinity

NADPHox_Plas 0 +infinity

NADPHox_Pero 0 +infinity

Chapter 3

56

Appendix 3.5.4 Genome-scale metabolic model of iHY3410 in ScrumPy

format. Can be found at: http://onlineli-

brary.wiley.com/doi/10.1111/tpj.13075/abstract

Appendix 3.5.5 Genome-scale metabolic model of iHY3410 in Excel for-

mat. Can be found at:

http://onlinelibrary.wiley.com/doi/10.1111/tpj.13075/abstract

Appendix 3.5.6 FBA and FVA results of dark, light and photorespiratory

scenarios for minimizing total flux. Can be found at: http://onlineli-

brary.wiley.com/doi/10.1111/tpj.13075/abstract

Chapter 4 In silico study of

photorespiratory metabolism

under drought conditions1

1 Part of this chapter has been published:

Yuan, H., Cheung, C.Y.M., Poolman, M.G., Hilbers, P.A.J. and van Riel,

N.A.W. (2016) A genome-scale metabolic network reconstruction of to-

mato (Solanum lycopersicum L.) and its application to photorespiratory

metabolism. Plant J., 85, 289–304.

Abstract

In this chapter, the reconstructed metabolic model of tomato described in Chapter 3

was used to investigate the metabolic characteristics for photorespiration and other

relevant metabolic processes under drought stress. Our results suggested that: (1)

the flux distributions through the mevalonate (MVA) pathway under drought were

distinct from that under normal conditions; (2) the changes in fluxes through core

metabolic pathways with varying flux ratio of RubisCO carboxylase to oxygenase

may contribute to the adaptive stress response of plants. In addition, previous studies

of reaction essentiality analysis for leaf metabolism were improved by including

potential alternative routes for compensating reaction knockouts. Altogether, the

genome-scale model gives valuable insights into the functional consequences of abi-

otic stresses.

Modelling photorespiratory metabolism under drought conditions

59

4.1 Introduction

The growth, yield and fruit quality of tomatoes can be affected by several environ-

mental factors, among which drought is a widespread abiotic stress, limiting the

growth and yield of crop plants (Haupt-Herting et al., 2001; Pinheiro and Chaves,

2011; Nakashima et al., 2014). Under drought conditions, photosynthesis is inhib-

ited mainly by stomatal limitation (Cornic, 2000; Flexas et al., 2007; Chaves et al.,

2009). Moreover, the occurrence of photorespiration dramatically increases under

drought stress as the leaf stomata are closed to prevent water loss, resulting in re-

duced CO2 concentration in leaves (Bauwe et al., 2012). Photorespiration is a

process in which O2 competes with CO2 to react with ribulose-1, 5-bisphosphate

(RuBP), thereby producing 3-phosphoglycerate (3-PGA) and 2-phosphoglycolate

(2-PG). Although this process represents a potential limitation on photosynthetic

efficiency and, ultimately, crop production, it has also been suggested that pho-

torespiration plays an important role in other metabolic processes such as energy

dissipation under stress conditions (Wingler et al., 2000).

Genome-scale metabolic modelling has been successfully used to predict met-

abolic states of plant metabolism (Williams et al., 2010; Simons et al., 2014). Thus

far, most modelling contributions have been made in the area of plant growth and

primary metabolism with very little work focusing on changes in metabolism in

response to abiotic stress such as drought. However, the effect of drought stress via

water deficit cannot be directly modelled using flux balance analysis (FBA). Con-

sidering the changes in biomass composition and growth rates caused by drought

(Boyer, 1982), the condition-specific growth rate that accounts for biomass accu-

mulation in drought conditions was introduced to mimics the effect of drought.

Simultaneously, the flux ratio of carboxylation to oxygenation of RubisCO (Vc/Vo)

was set to a value of 1 to represent drought stress (Jordan and Ogren, 1984).

In this chapter, the tomato model described in Chapter 3 was used to investigate

the cellular metabolic characteristics, in particular the interplay of photorespiration

with other pathways, under drought stress by identifying changes in reaction fluxes.

The model predicted that the flux distributions through the mevalonate (MVA) path-

way under drought conditions significantly differed from that under normal

conditions, which gave hints to possible adaptive metabolic responses to drought

stresses.

Chapter 4

60

4.2 Methods

4.2.1 Model analysis

The tomato model was analysed using the same methods as described in Chapter 3,

section 3.2.3. Photorespiratory metabolism was simulated by additionally setting the

flux ratio of carboxygenic to oxygenic of RubisCO (Vc/Vo), which is 3 under am-

bient conditions, and 1 under drought stressed conditions. For the simulations of

drought conditions, the biomass outputs were constrained using the drought-specific

growth rate (Appendix 3.5.1). Other constraints for simulating the photorespiratory

scenario are the same as described in Chapter 3, section 3.2.3.

4.2.2 Deletion simulation

Large-scale metabolic models serve as platforms for rapidly predicting growth phe-

notypes in response to reaction knockouts in silico. In order to simulate reaction

deletions, the flux through the corresponding reaction was constrained to zero in

addition to the constraints on biomass production, ATP maintenance, nitrogen input

and Vc/Vo. With the constraints of biomass synthesis, a reaction was categorized as

essential if its deletion resulted in no feasible steady-state solution. In contrast, re-

actions were defined as non-essential if a feasible solution could be found.

4.3 Results and Discussion

As previously mentioned, drought is a common abiotic stress that limits tomato

growth, yield and fruit quality. To elucidate the metabolic response of tomato plants

to water-limited conditions, the drought stress was modelled using our reconstructed

model by incorporating the drought-specific growth rate of tomato (2.84 mg g-1 DW

h-1) based on the work of Sánchez-Rodríguez et al. (2010) while constraining the

ratio of Vc/Vo to a value of 1 (i.e. Vc/Vo=1), which results from the stomata closure

in response to drought stress that leads to a reduced CO2 concentration in leaves

(Bauwe et al., 2012; Lakshmanan et al., 2013b). The metabolic characteristics of

drought conditions was then compared with that of normal conditions with a growth

rate of 3.75 mg g-1 DW h-1 (Sánchez-Rodríguez et al., 2010), and the ratio of Vc/Vo

was set to 3 (i.e. Vc/Vo=3). In total, 376 and 371 reactions were found to be active

under drought and normal conditions, respectively. A set of 358 active reactions

were common between the two flux solutions. It is expected that the photon uptake

Modelling photorespiratory metabolism under drought conditions

61

rate under stressed conditions (13.32 mmol g-1 DW h-1) was higher than that under

normal conditions (10.89 mmol g-1 DW h-1) as more energy is required for the recy-

cling of higher amounts of 2-PG produced under drought stress.

4.3.1 Differences in metabolically active reactions between normal

and drought conditions

Sets of reactions which were only used either in normal or in drought conditions

were identified according to results from FVA (Appendix 4.5.1), and these reactions

were investigated in more detail. A total of 10 reactions were identified to carry non-

zero flux only under normal conditions, which implies these reactions were always

inactive under drought conditions. Among these reactions, 8 reactions (EC 2.2.1.7;

1.1.1.267; 2.7.7.60; 2.7.1.148; 4.6.1.12; 1.17.1.2; 1.17.7.1) are involved in the 2C-

methyl-D-erythritol-4-phosphate pathway or MEP pathway (Table 4.1). On the

other hand, 14 reactions were found to be active under drought conditions, of which

7 (EC 6.2.1.16; 1.1.1.34; 2.7.1.36; 2.7.4.2; 4.1.1.33; 2.3.1.9; 4.1.3.4) belong to

MVA pathway, whilst, 3 are the related metabolite transporters (‘CPD-

499_Pero_tx’, ‘DELTA3-IPP_Pero_tx’ and ‘DELTA3-IPP_Plas_tx’). In plants,

both the MEP pathway and the MVA pathway are predominantly used to synthesize

isopentenyl pyrophosphate (IPP) and dimethylallyl pyrophosphate (DMAPP),

which can be further utilized for the biosynthesis of isoprenoids (Vranová et al.,

2013). Isoprenoids, also called terpenoids, are required for the production of a vari-

ety of important compounds such as chlorophylls, carotenoids, and plant hormones.

The MEP pathway utilizes pyruvate and D-glyceraldehyde-3-phosphate (GAP) for

IPP production, which takes place in the plastid. Isoprenoid biosynthesis through

the MEP pathway depends on the availability of reduced carbon, ATP and NADPH,

which are contributed by photosynthesis in unstressed plants (Loreto and Schnitzler,

2010). In contrast, the MVA pathway in the cytosol and the peroxisome uses acetyl-

CoA as the precursor for IPP synthesis. Ultimately, IPP generated either via the

MEP pathway or the MVA pathway was used for the biosynthesis of chlorophylls

in our model (Figure 4.1).

Chapter 4

62

Table 4.1 List of unique reactions carrying non-zero flux under either normal or drought conditions.

Vc/Vo=3

Identifier Description Identifier Description

DXS-RXN_Plas EC 2.2.1.7; MEP path-

way 1.1.1.34-RXN_Cyto

EC 1.1.1.34; MVA

pathway

DXPREDISOM-

RXN_Plas

EC 1.1.1.267; MEP

pathway

MEVALONATE-KINASE-

RXN_Cyto

EC 2.7.1.36; MVA

pathway

2.7.7.60-RXN_Plas EC 2.7.7.60; MEP path-

way

PHOSPHOMEVALONATE-

KINASE-RXN_Pero

EC 2.7.4.2 ; MVA

pathway

2.7.1.148-

RXN_Plas

EC 2.7.1.148; MEP

pathway IPPISOM-RXN_Plas

EC 5.3.3.2;

MVA/MEP pathway

RXN0-302_Plas EC 4.6.1.12; MEP path-

way

DIPHOSPHOMEVALONTE-

DECARBOXYLASE-RXN_Pero

EC 4.1.1.33 ; MVA

pathway

RXN0-884-

(NADP)_Plas

EC 1.17.1.2; MEP path-

way

ACETYL-COA-

ACETYLTRANSFER-RXN_Cyto

EC 2.3.1.9; MVA

pathway

RXN0-882_Plas EC 1.17.7.1; MEP path-

way

HYDROXYMETHYLGLUTARYL-

COA-LYASE-RXN_Cyto

EC 4.1.3.4 ; MVA

pathway

ISPH2-RXN-

(NADP)_Plas

EC 1.17.1.2; MEP path-

way CPD-499_Pero_tx

peroxisomal phos-

phomevalonate

transporter

CDPKIN-

RXN_Plas

EC 2.7.4.6; CMP phos-

phorylation pathway DELTA3-IPP_Plas_tx

plastidic isopentenyl

pyrophosphate trans-

porter

CMPKI-RXN_Plas EC 2.7.4.14; CMP phos-

phorylation pathway DELTA3-IPP_Pero_tx

peroxisomal isopen-

tenyl pyrophosphate

transporter

Vc/Vo=1 CO2_Pero_tx peroxisomal CO2

transporter

ACETOACETATE-

-COA-LIGASE-

RXN_Cyto

EC 6.2.1.16; MVA path-

way

ATP_ADP_Pero_tx peroxisomal ATP

transporter

ATP-CITRATE-

PRO-S--LYASE-

RXN_Cyto

EC 2.3.3.8; acetyl-CoA

biosynthesis pathway

Modelling photorespiratory metabolism under drought conditions

63

Figure 4.1 Subcellular compartmentalization of the MVA and MEP pathways and ABA biosynthesis. Red lines

indicate the corresponding reactions are only active under normal conditions (Vc/Vo=3); blue lines indicate reac-

tions are only active under drought conditions (Vc/Vo=1); black solid lines indicate reactions are active both under

normal and drought conditions; dashed lines indicate those reactions are not active in the prediction. 3-PGA, 3-

phosphoglycerate; PEP, phosphoenol pyruvate; OAA, oxalacetic acid; GAP, glyceraldehyde-3-phosphate; MEP,

2C-methyl-D-erythritol-4-phosphate; IPP, isopentenyl pyrophosphate; MVA, mevalonate; MVA-5P, mevalonate-

5-phosphate; ABA, abscisic acid.

Compartmentalization of IPP-generating pathways would enable plants to op-

timize isoprenoid biosynthesis and regulation according to fixed carbon, available

ATP and reducing power in diverse environmental conditions. Previous studies have

shown that chlorophylls and carotenoid-related pigments are derived from the plas-

tidic MEP pathway (Lichtenthaler et al., 1997), while MVA-derived precursors can

be transported into the plastid and used for the isoprenoids biosynthesis in the plastid

(Wang et al., 2003; Rodríguez-Concepción et al., 2004). Under normal conditions,

our model predicted that IPP was synthesised via the chloroplastic MEP pathway,

which supports previous findings that isoprenoid is produced primarily using carbon

skeletons directly derived from the Calvin cycle via the photosynthesis-dependent

Chapter 4

64

MEP pathway under stress-free conditions (Affek and Yakir, 2002; Fortunati et al.,

2008). However, under drought conditions, our model predicted that the MVA path-

way predominated over the MEP pathway for producing IPP for chlorophyll

biosynthesis. The precursor for IPP biosynthesis through the MVA pathway is cy-

tosolic acetyl-CoA. A detailed investigation of the flux solution under drought

conditions revealed a metabolic route that produces acetyl-CoA from primary pho-

tosynthates via phosphoenolpyruvate carboxylase (PEPC) and threonine aldolase

(Figure 4.1). For each molecule of 3-PGA converted into oxaloacetate (OAA) via

phosphoenolpyruvate (PEP), one molecule of CO2 is fixed by PEPC. OAA is then

used as carbon skeleton to form threonine. Threonine aldolase splits threonine into

acetaldehyde, which is converted into acetyl-CoA after oxidation to acetate, and

glycine, which is recycled through the photorespiratory pathway to produce 3-PGA.

For each molecule of glycine recycled, 0.5 molecule of CO2 is released and 0.5

molecule of 3-PGA is produced. In summary, the carbon of acetyl-CoA comes from

0.5 molecule of 3-PGA and 0.5 molecule of CO2 of which one molecule is fixed by

PEPC and 0.5 molecule is released from the photorespiratory pathway. While this

route appears to be inefficient in normal conditions, it could become favorable under

drought condition when the leaf internal CO2 concentration is so low that the cost

of fixing CO2 by Rubisco coupled with a high level of photorespiration becomes

higher than the cost of the PEPC-threonine aldolase route in which some of the car-

bon for IPP synthesis is fixed by PEPC. The PEPC-threonine aldolase route has also

been proposed as a possible route for producing acetyl-CoA for citrate synthesis in

leaves growing under constant illumination (Cheung et al., 2014).

It is evident that drought stress leads to a substantial reduction in the rate of

photosynthetic CO2 assimilation, resulting in limited photosynthate accumulation

and an increased supply of photosynthetic energy and reducing power to non-pho-

tosynthetic carbon reduction sinks such as photorespiration. It has been proposed

that the MEP pathway competes with other sinks such as photorespiration for the

reducing power, which is not invested in carbon assimilation for primary metabo-

lism (Dani et al., 2014; Morfopoulos et al., 2014). However, the reducing power

requirement by the MEP pathway is relatively small compared to that of photorespi-

ration (Sharkey et al., 2008; Dani et al., 2014). When photosynthate is limiting,

alternate carbon sources such as intermediates from the MVA pathway and pho-

torespiration could directly contribute to isoprenoid production in plants (Flügge

and Gao, 2005; Jardine et al., 2014). Isoprenoid emissions can even continue in the

complete absence of photosynthesis under severe drought conditions (Pegoraro et

al., 2004; Fortunati et al., 2008). Furthermore, it has been suggested that genes en-

coding in the MVA pathway are activated and highly expressed in response to

Modelling photorespiratory metabolism under drought conditions

65

oxidative stress (Vranová et al., 2013), which also supports our model prediction

that the MVA pathway was used for IPP biosynthesis under drought conditions.

Given that minimization of total flux was used as objective function in this

study, it can be inferred that the MEP pathway is more efficient in terms of amount

of flux for the biosynthesis of IPP than the MVA pathway under normal conditions,

and vice versa for drought conditions. To confirm this, four scenarios were simu-

lated, in which IPP can only be produced either via MEP pathway or MVA pathway.

The 4 scenarios were denoted as MEP:Normal, MVA:Normal, MEP:Drought, and

MVA:Drought. As expected, the sum of total flux for MEP:Normal scenario was

lower than that for MVA:Normal scenario, which were 65.85 and 65.95 mmol g-1

DW h-1, respectively. Likewise, the total flux of MVA:Drought scenario was lower

than that of MEP:Drought scenario, which were 77.86 and 77.89 mmol g-1 DW h-1,

respectively. Regarding the photon influx, MEP:Normal scenario required less pho-

tons than that for MVA:Normal scenario (10.89 and 10.90 mmol g-1 DW h-1,

respectively), despite a higher total Rubisco flux with the MEP pathway (0.320 and

0.318 mmol g-1 DW h-1, respectively). By contrast, MEP:Drought scenario required

more photons than that for MVA:Drought scenario (13.34 and 13.32 mmol g-1 DW

h-1, respectively) because the photon requirement for the net fixation of CO2 was

higher under drought conditions (total Rubisco flux was 0.604 and 0.600 mmol g-1

DW h-1 for MEP:Drought scenario and MVA:Drought scenario, respectively).

It is noteworthy that abscisic acid (ABA), belonging to isoprenoids, derives

from the intermediate IPP (Figure 4.1). A wealth of evidence supports the hypothe-

sis that ABA is a regulator triggering plant responses to numerous abiotic stresses

such as drought (Cornish and Zeevaart, 1985; Harris and Outlaw, 1991). During

drought stress, the synthesis of ABA was induced rapidly, whereas the increased

amounts of ABA promoted stomatal closure to reduce water loss (Bray, 1988). ABA

in higher plants is derived from carotenoid precursors through an oxidative cleavage

reaction in plastid, followed by a two-step conversion of the xanthoxin to ABA via

ABA-aldehyde (Xiong and Zhu, 2003). However, because carotenoids are not in the

output constraints of the model, this ABA biosynthetic route was not active in our

model prediction, whilst, the observations that reactions in the MVA pathway were

activated for biomass synthesis may relate to an increase in ABA biosynthesis in

response to drought stress.

As sessile organism, plants have evolved several mechanisms that allow them

to cope with drought stress, for example, increased accumulation of secondary me-

tabolites such as phenols and isoprenoids (Wilhelm and Selmar, 2011). It has been

documented that isoprenoids can provide protection from abiotic stresses, e.g.

through quenching the accumulation of reactive oxygen species (ROS), which result

Chapter 4

66

in oxidative damage (Loreto and Velikova, 2001). One of the inevitable conse-

quences caused by drought stress is the over accumulation of ROS (Smirnoff, 1993).

The switch from the MEP pathway to the MVA pathway could also be due to iso-

prenoid production, which allows plants to protect against the deleterious effects

caused by drought stress.

To test if the operation of the MVA pathway under drought conditions is related

to the choice of the objective function, flux solutions were calculated using minimi-

zation of photon uptake as the objective function as well (Appendix 4.5.2).

Minimization of photon uptake gave similar flux solutions to those from minimiza-

tion of total flux in regard to IPP biosynthesis, which strengthens our hypothesis

that IPP is produced via the MEP pathway under normal conditions, but switched to

the MVA pathway in drought conditions.

4.3.2 Effect of drought stress on the core metabolic pathways

Photorespiratory metabolism affects many parts of cellular metabolism, including

the Calvin cycle, the TCA cycle as well as nitrogen assimilation (Foyer et al., 2009).

To understand the coordinated function of photorespiration with other metabolic

pathways, the change in fluxes between normal and drought conditions for reactions

involved in several important metabolic processes was investigated (Table 4.2).

With exposure to drought stress, sharp increases in fluxes through the Calvin cycle

and photorespiration were observed to support the biomass synthesis. Growth re-

duction under water-deficit conditions has been well documented (Sánchez-

Rodríguez et al., 2010; Pinheiro and Chaves, 2011).

In the model flux solutions, reactions involved in ammonia assimilation (the

mitochondrial GS and Fd-GOGAT) showed significant differences in fluxes be-

tween normal and drought conditions. Large amounts of ammonia released during

photorespiration via GDC in the mitochondria, were re-assimilated via mitochon-

drial GS. Similar to the flux distribution under non-photorespiratory light conditions

(Chapter 3, Figure 3.3b), reactions through the OPP pathway were inactive, and the

TCA cycle exhibited a noncyclic flux mode under stressed conditions. Haupt-Hert-

ing et al. (2001) have showed that the rate of mitochondrial respiration in the light

is reduced under drought stress in tomato mature leaves, while, a growing leaf cell

was modellied. Though a growing body of evidence from tomato and Arabidopsis

suggests that the activity of the TCA cycle was considerably higher in the light than

reported by Tcherkez et al. (2005), the degree of inhibition of the TCA cycle in the

light is still controversial (Araújo et al., 2012). Compared with other TCA reactions,

the reaction catalyzed by malate dehydrogenase (MDH) carried a relatively high

Modelling photorespiratory metabolism under drought conditions

67

flux in both normal and drought conditions in our prediction. This could be because

MDH functions in redox perturbation but also acts as a component of the malate-

aspartate and malate-oxaloacetate shuttles for the exchange of reducing equivalents

between the mitochondria and the cytosol.

Table 4.2 Predicted flux levels (µmol/g DW/h) of the main metabolic reactions under normal (Vc/Vo=3) and

drought conditions (Vc/Vo=1).

Enzyme name EC Biochemical reaction Vc/Vo=3 Vc/Vo=1

Calvin Cycle

RubisCO carboxylase 4.1.1.39 RuBP +CO2 -> 2 3-PGA 239.67 300.11

PGK 2.7.2.3 3-PGA + ATP <-> DPG + ADP 576.23 1,037.85

GAPDH 1.2.1.13 DPG + NADPH -> GAP + NADP + Pi 576.23 1,038.23

TPI 5.3.1.1 GAP <-> DHAP 232.72 404.30

Aldolase 4.1.2.13 GAP + DHAP <-> FBP -110.01 -202.65

FBPase 3.1.3.11 FBP -> F6P + Pi 110.01 202.65

TKT 2.2.1.1 GAP + F6P <-> E4P + X5P -110.01 -202.65

Aldolase 4.1.2.13 DHAP + E4P -> SBP 108.69 201.65

SBPase 3.1.3.37 SBP -> S7P + Pi 108.69 201.65

TKT 2.2.1.1 GAP + S7P <-> R5P + X5P -108.69 -201.65

RPE 5.1.3.1 X5P <-> Ru5P -218.75 -404.34

RPI 5.3.1.6 R5P <-> Ru5P 101.07 195.88

PRK 2.7.1.19 Ru5P + ATP -> RuBP + ADP 319.82 600.22

Photorespiration

RubisCO oxygenase 4.1.1.39 RuBP + O2 -> 3-PGA + 2-PG 79.96 300.11

PLGP 3.1.3.18 2-PG -> Glt + Pi 79.96 300.11

GOX 1.1.3.15 Glt + O2 -> Gox + H2O2 79.96 300.11

CAT 1.11.1.6 2 H2O2-> 2 H2O + O2 39.97 150.05

GTR 2.6.1.4 Glu + Gox -> Gly + 2-OG -46.54 -153.95

AGT 2.6.1.44 Ala + Gox -> Gly + Pyr - -

GDC

1.4.4.2;

2.1.2.10;

1.8.1.4

Gly + THF + NAD+ -> CO2 + NH3 +NADH + M-THF 45.55 155.35

SHMT 2.1.2.1 Gly + M-THF <-> Ser + THF - -

SGT 2.6.1.45 Ser + Gox -> OH-PYR + Gly 33.51 146.23

HPR 1.1.1.29 OH-PYR + NADH -> Glycerate + NAD 33.51 146.23

GLYK 2.7.1.31 Glycerate + ATP -> 3-PGA + ADP 33.51 146.23

TCA

PDH - Pyr + NAD+ + CoA -> AcCoA + NADH + H+ + CO2 - -

CIT 2.3.3.1 OAA + AcCoA + H2O -> Cit + CoA - -

Aconitase 4.2.1.3 Cit <-> IsoCit 3.77 2.84

IDH 1.1.1.41 IsoCit + NAD+ -> 2-OG + NADH + H+ +CO2 3.77 2.84

OGDH

1.2.4.2;

1.8.1.4

2-OG + NAD+ + CoA -> SucCoA + NADH + H+ + CO2 - -

STK 6.2.1.5 SucCoA + Pi + ADP + H+ -> Suc + ATP +CoA - -

Chapter 4

68

Complex II

1.3.99.1;

1.3.5.1 Suc + Q -> Fum + QH2 - -

Fumarase 4.2.1.2 Fum + H2O -> Mal - -

MDH 1.1.1.37 Mal + NAD+ <-> OAA + NADH + H+ 1,969.48 1,921.19

OPPP

G6PD 1.1.1.49 G6P + NADP+ -> 6PGL + NADPH + H+ - -

6PGL 3.1.1.31 6PGL + H2O -> 6PG - -

6PGD 1.1.1.44 6PG + NAD(P)+ <-> Ru5P + NAD(P)H + CO2 + H+ - -

ETC

Complex I 1.6.5.3 NADH + Q + 5 H+ -> NAD+ + QH2 + 4 H+ 2,018.80 2,079.37

Complex III 1.10.2.2 QH2 + 2Cytox + 2 H+ -> Q + 2Cytred + 4 H+ 2,021.11 2,081.13

Complex IV 1.9.3.1 4Cytred + O2 + 8 H+ -> 4 Cytox + 2H2O + 4 H+ 1,010.59 1,040.59

Complex V 3.6.3.14 ADP + Pi + 4 H+ <-> ATP + H2O + 4 H+ 1,557.50 1,612.28

Nitrogen assimilation

NiR 1.7.7.1 NO3 + NADH -> NO2 + NAD+ 17.62 13.34

NR 1.7.1.2 Fdred + NO2 -> Fdox + NH4+ 17.62 13.34

GS_Plas 6.3.1.2 NH3 + Glu + ATP -> Gln + ADP + Pi + H+

35.10 26.56

GS_Mito 45.55 155.35

GOGAT 1.4.1.14 Gln + 2-OG + NADH + H+ -> 2 Glu + NAD+ - -

1.4.7.1 Gln + 2-OG + 2Fdred + H+ -> 2 Glu +2Fdox 70.01 173.86

‘-’ indicates no flux. ‘GS_Plas’ indicates the plastidic GS, ‘GS_Mito’ indicates the mitochondrial GS.

4.3.3 Evaluation of reaction deletions

A genome-scale metabolic model is a powerful tool to assess the metabolic capabil-

ities of an organism. Aiming to examine the phenotypes of genetic manipulations

with the model, a reaction essentiality analysis for normal (Vc/Vo=3) and drought

stressed conditions (Vc/Vo=1) was performed. To investigate the essentiality of

each reaction, an additional constraint was added by setting each reaction except the

biomass transporters to carry zero flux in addition to the general constraints de-

scribed in section 4.2.2.

In total, 146 and 149 lethal reaction knockouts were identified for normal and

drought conditions, respectively, indicating that these reactions must carry non-zero

fluxes to support biomass synthesis under the simulated conditions (Appendix

4.5.3). All 146 essential reactions identified for normal conditions were also indis-

pensable for drought conditions. With respect to the ability to produce biomass,

approximately 93% of the reactions in the model were non-essential, reflecting that

the metabolic network was robust to random knockout mutations. A number of key

enzymes mainly from the Calvin cycle and the photorespiratory pathway were se-

lected, and their essentiality in models of tomato, rice, Arabidopsis and maize were

Modelling photorespiratory metabolism under drought conditions

69

compared for further verification (Table 4.3). In general, these predictions are sup-

ported by the existing experimental evidence available in other plants such as

Arabidopsis and barley.

Although a few of the essential reactions are common across all 4 plants, such

as RubisCO, ribulose-5-phosphate epimerase and PGLP, our predictions showed

significant difference from that of rice, Arabidopsis and maize (Table 4.3). The es-

sential reactions of iHY3410 listed in Table 4.3 were identified as essential in rice

as well (Lakshmanan et al., 2013b). However, most of photorespiratory reactions

including glutamate:glyoxylate aminotransferase (GGT), serine:glyoxylate ami-

notransferase (SGT), SHMT, hydroxypyruvate reductase (HPR), glycerate kinase

(GLYK) and catalase (CAT) can be bypassed by other isoforms or pathways in

iHY3410, but were regarded as essential in the core rice model (Lakshmanan et al.,

2013b). Importantly, the predictions of essentiality are critically dependent on the

coverage and completeness of the model. In the case of rice, these simulation results

were obtained from a core metabolic network with 417 reactions and only three

subcellular compartments (cytosol, chloroplast, and mitochondrion) without the pe-

roxisome. In comparison, our tomato model was not only constructed at the genome-

scale level but also more compartmentalized. It is likely that the addition of com-

partmentation in a model will give raise to parallel and alternative pathways

(Sweetlove and Ratcliffe, 2011). For example, glyoxylate oxidized from glycolate

by GOX is subsequently metabolized by aminotransferases which redistributes the

nitrogen to a range of amino acids, which then are used for synthesis of the other

amino acids. The assignment of the reactions involved in photorespiratory metabo-

lism in our tomato model and the rice model were illustrated in Figure 3.4 (Chapter

3) and Appendix 4.5.4, respectively, and were summarized in Appendix 4.5.5.

Though the directionality of these reactions was consistent between the two models,

the assigned compartments were different for most of the reactions. In the peroxi-

some of iHY3410, there were three aminotransferases involved in the redistribution

of amino group in the photorespiratory pathway, namely SGT, GGT and ala-

nine:glyoxylate aminotransferase (AGT). As a result, unlike other photorespiratory

mutants (e.g. PGLP), mutants defective in SGT, GGT or AGT activity did not result

in a lethal phenotype in our predictions from iHY3410. By contrast, SGT was clas-

sified as essential in the core rice model in which AGT was not included. Likewise,

our simulations showed that knockout of RPI (Vc/Vo=3) and HPR can be comple-

mented by alternative routes (Figure 4.2 and Figure 4.3, respectively). These

examples illustrate that the application of essentiality analysis on a core metabolic

network could give an overestimation of essential reactions as potential alternative

Chapter 4

70

Table 4.3 Comparisons of essential reactions with other models. ‘+’, essential; ‘-’, non-essential; ‘Unknown’, not reported; ‘#’, essential only under stressed conditions.

Enzyme

EC Pathway iHY3410 Rice Arabidopsis Maize

References

Lakshmanan et al. (2013a) Wang et al., 2012

RubisCO 4.1.1.39 Calvin cycle + + + + Quick et al. (1991)

PRK 2.7.1.19 Calvin cycle + + + + Moll and Levine (1970)

RPE 5.1.3.1 Calvin cycle/PPP + + + +

RPI 5.3.1.6 Calvin cycle/PPP # + + + Xiong et al. (2009)

TKT 2.2.1.1 Calvin cycle/PPP + + + +

SBPase 3.1.3.37 Calvin cycle - + Unknown Unknown

PEPC 4.1.1.31 C4 photosynthesis - - - + Rademacher et al. (2002)

PPDK 2.7.9.1 C4 photosynthesis - Unknown - - Mechin et al. (2007)

PLGP 3.1.3.18 Photorespiration + + - - Schwarte and Bauwe (2007)

GOX 1.1.3.15 Photorespiration + + + + Xu et al. (2009)

GGT 2.6.1.4 Photorespiration - Unknown Unknown Unknown

SGT 2.6.1.45 Photorespiration - + Unknown Unknown Liepman and Olsen (2001)

GDC

1.4.4.2;

2.1.2.10;

1.8.1.4

Photorespiration # + Unknown Unknown Engel et al. (2007)

SHMT 2.1.2.1 Photorespiration - # Unknown Unknown

HPR 1.1.1.29 Photorespiration - Unknown Unknown Unknown

GLYK 2.7.1.31 Photorespiration - + Unknown Unknown

CAT 1.11.1.6 Photorespiration - + Unknown Unknown

GS 6.3.1.2 GS/GOGAT cycle - + Unknown Unknown

Fd-GOGAT 1.4.7.1 GS/GOGAT cycle - + Unknown Unknown

GLBE 2.4.1.18 Starch biosynthesis + + + + Hernández et al. (2008)

TAL 2.2.1.2 PPP - Unknown - -

Aconitase 4.2.1.3 TCA - Unknown + + Carrari et al. (2003)

Modelling photorespiratory metabolism under drought condtions

71

routes that may compensate the reaction deletion were not included in the core

model.

Figure 4.2 The bypass of RPI knockout under normal conditions (Vc/Vo=3) in iHY3410. Grey line indicates the

reaction is blocked.

The model predictions on reaction essentiality may also differ depending on

the conditions simulated. For example, in Arabidopsis and maize, 1,4-alpha-glucan

branching enzyme (EC 2.4.1.18) was essential when optimizing biomass produc-

tion, while in the same networks, its knockout had no lethal effect regarding CO2

fixation (Wang et al., 2012). A handful of photorespiratory mutants in C3 species

have been generated, notably Arabidopsis and barley, but a very limited set has been

generated for tomato (Foyer et al., 2009). Thus, the majority of the in silico pho-

torespiratory mutants predicted from our tomato model remain to be validated

experimentally.

Chapter 4

72

Figure 4.3 The bypass of HPR knockout in iHY3410. Numbered reactions are catalyzed by the following enzymes:

1, serine dehydratase (EC 4.3.1.17); 2, pyruvate phosphate dikinase (EC 2.7.9.1); 3, phosphopyruvate hydratase

(EC 4.2.1.11); 4, phosphoglycerate mutase (EC 5.4.2.12).

4.4 Conclusions

The reconstructed genome-scale metabolic model of tomato, iHY3410 allowed us

to simulate the metabolic behavior under various environmental conditions,

including normal and drought conditions using flux balance analysis. Our model

simulations predicted that the flux distributions through the MVA pathway under

drought stressed conditions were distinct from that under normal conditions. In

addition, the change in fluxes through the core metabolic pathways between normal

and drought conditions reflected the plant adaptive response to drought stress. The

application of genome-scale networks to the abiotic stresses in plants is still in its

infancy, and the presented tomato model can serve as a starting point for studying

the functional consequences of abiotic stresses. While iHY3410 was used for

analysing the behaviour of leaves in this study, it could also be used for exploring

the metabolic behaviour of tomato fruit in future studies.

4.5 Appendix

Appendix 4.5.1 FBA and FVA results of photorespiratory scenarios in nor-

mal and drought conditions for minimizing total flux. Can be found at:

http://onlinelibrary.wiley.com/doi/10.1111/tpj.13075/abstract

Modelling photorespiratory metabolism under drought conditions

73

Appendix 4.5.2 FBA and FVA results of reactions involved in the MEP and MVA pathways in normal and drought

conditions for minimizing photon uptake.

Reaction ID Classification Vc/Vo=3 Vc/Vo=1

FBA FVA

FBA FVA

Min Max Min Max

ISPH2-RXN-(NAD)_Plas MEP pathway 0 0.0 0.9532 0 0.0 0.0

RXN0-884-(NADP)_Plas MEP pathway 0.2383 0.0 0.9532 0 0.0 0.0

RXN0-302_Plas MEP pathway 0.9532 0.953199999988 0.9532 0 0.0 0.0

RXN0-884-(NAD)_Plas MEP pathway 0 0.0 0.9532 0 0.0 0.0

2.7.1.148-RXN_Plas MEP pathway 0.9532 0.953199999988 0.9532 0 0.0 0.0

DXPREDISOM-RXN_Plas MEP pathway 0.9532 0.953199999988 0.9532 0 0.0 0.0

2.7.7.60-RXN_Plas MEP pathway 0.9532 0.953199999988 0.9532 0 0.0 0.0

ISPH2-RXN-(NADP)_Plas MEP pathway 0.7149 0.0 0.9532 0 0.0 0.0

DXS-RXN_Plas MEP pathway 0.9532 0.953199999988 0.9532 0 0.0 0.0

RXN0-882_Plas MEP pathway 0.9532 0.953199999988 0.9532 0 0.0 0.0

IPPISOM-RXN_Plas

MEP pathway;

MVA pathway 0 -0.7149 0.2383 0.18 0.180000000009 0.18

ACETOACETATE--COA-LIGASE-RXN_Cyto MVA pathway 0 0.0 0.0 -0.72 -0.720000000001 -0.720000000349

DIPHOSPHOMEVALONTE-DECARBOXYLASE-

RXN_Pero MVA pathway 0 0.0 0.0 0.72 0.720000000028 0.720000000001

1.1.1.34-RXN_Cyto MVA pathway 0 0.0 0.0 0.72 0.0 0.720000000001

ACETYL-COA-ACETYLTRANSFER-RXN_Cyto MVA pathway 0 0.0 0.0 0.72 0.720000000002 0.720000000001

CPD-499_Pero_tx MVA pathway 0 0.0 0.0 -0.72 -0.720000000001 -0.720000000028

DELTA3-IPP_Pero_tx MVA pathway 0 -inf inf 0.72 -inf inf

MEVALONATE-KINASE-RXN_Cyto MVA pathway 0 0.0 0.0 0.72 0.720000000028 0.720000000001

DELTA3-IPP_Plas_tx MVA pathway 0 0.0 0.0 0.72 0.720000000028 0.719999999999

HYDROXYMETHYLGLUTARYL-COA-LYASE-

RXN_Cyto MVA pathway 0 0.0 0.0 -0.72 -0.720000000001 -0.720000000349

PHOSPHOMEVALONATE-KINASE-RXN_Pero MVA pathway 0 0.0 0.0 0.72 0.720000000028 0.720000000001

Chapter 4

74

Appendix 4.5.3 List of essential reactions of tomato leaves for normal and

drought conditions. Can be found at: http://onlineli-

brary.wiley.com/doi/10.1111/tpj.13075/abstract

Appendix 4.5.4 Assignment of the reactions involved in the photorespira-

tory metabolism in the rice model by Lakshmanan et al. (2013a).

Numbered reactions are catalyzed by the following enzymes: 1, RubisCO oxygenase; 2, PLGP; 3, GOX; 4, CAT;

5, GGT; 6, GDC; 7, SHMT; 8, GS; 9, SGT; 10, HPR; 11, GLYK; 12, GOGAT.

Modelling photorespiratory metabolism under drought conditions

75

Appendix 4.5.5 Comparisons of photorespiratory reactions between iHY3410 and a rice core model.

Reaction

Name EC Reaction Stoichiometry

iHY3410 Rice (Lakshmanan et al., 2013a)

Reaction ID Compartment Directionality Reaction ID Compartment Directionality

RubisCO

oxygenase 4.1.1.39 RuBP + O2 -> 3-PGA + 2-PG RXN-961_Plas plastid -> RBCS_O_p_ plastid ->

PLGP 3.1.3.18 2-PG -> Glt + Pi GPH-RXN_Plas plastid -> PGP_p_ plastid ->

GOX 1.1.3.15 Glt + O2 -> Gox + H2O2 RXN-969_Pero peroxisome -> GOX_c_ cytosol ->

CAT 1.11.1.6 2 H2O2-> 2 H2O + O2 CATAL-RXN_Pero peroxisome -> CAT_c_ cytosol ->

GGT 2.6.1.4 Glu + Gox -> Gly + 2-OG GLYCINE-AMINOTRANSFERASE-

RXN_Pero peroxisome -> GGAT_c_ cytosol ->

AGT 2.6.1.44 Ala + Gox -> Gly + Pyr ALANINE--GLYOXYLATE-

AMINOTRANSFERASE-RXN_Pero peroxisome -> Not included Not included Not included

GDC

1.4.4.2;

2.1.2.10;

1.8.1.4

Gly + THF + NAD -> CO2 + NH3

+NADH + M-THF GCVMULTI-RXN_Mito mitochondria -> GDC_m_ mitochondria ->

SHMT 2.1.2.1 Gly + M-THF <-> Ser + THF GLYOHMETRANS-RXN_Mito mitochondria <=> SHM1_m_ mitochondria <=>

SGT 2.6.1.45 Ser + Gox -> OH-PYR + Gly SERINE-GLYOXYLATE-

AMINOTRANSFERASE-RXN_Pero peroxisome -> SGAT_c_ cytosol ->

HPR 1.1.1.29 OH-PYR + NADH -> Glycerate + NAD GLYCERATE-DEHYDROGENASE-

RXN_Pero peroxisome -> HPR_c_ cytosol ->

GLYK 2.7.1.31 Glycerate + ATP -> 3-PGA + ADP GLY3KIN-RXN_Plas plastid -> GLYK_p_ plastid ->

GS 6.3.1.2 NH3 + Glu + ATP -> Gln + ADP + Pi +

H+ GLUTAMINESYN-RXN_Mito mitochondria -> GLN2_p_ plastid ->

Chapter 5 Modelling the effects of

light intensity and quality on

light-driven metabolism of

tomato1

1 Part of this chapter is described in:

Yuan, H., Cheung, C.Y.M., Hilbers, P.A.J. and van Riel, N.A.W. Modelling the

effects of light intensity and quality on the light-driven metabolism of to-

mato. In preparation

Abstract

The light environment of plants can highly vary in intensity and quality, both of

which markedly impact plant growth and development. Such influences have been

studied extensively, but often independently. Thus, very little is known about the

influence of changing light intensity on the cellular growth and photosynthetic pro-

cess of plants grown under different light sources. In this chapter, the cellular growth

and its metabolic properties in response to changing light intensity and quality were

studied using a light-specific metabolic model of tomato. The in silico analysis of

the light-driven process was conducted under low light (LL, 50 µmol/m2/s) and high

light intensity (HL, 700 µmol/m2/s) for seven different light sources, respectively.

The model successfully described the different responses of leaf cells to quantitative

and qualitative changes in light. The in-depth investigation of energy balancing

highlights the importance of mitochondrial respiration in regulating the

ATP/NADPH ratio for optimal growth. Our model predictions showed that, in the

LL state of yellow light, aldehyde dehydrogenase (ALDH) may represent an im-

portant metabolic interaction between plastid and other intracellular organelles. This

may help in a better understanding of photosynthetic response and adaption to the

changing light condition.

Modelling the effects of light intensity and quality

79

5.1 Introduction

Plants live in fluctuating light conditions, in which both light intensity and light

quality change by several orders of magnitude during the day and throughout the

year (Eaton-Rye et al., 2011). Such variations markedly affect growth, development

and yield. Regarding the effect of light quality, numerous publications have been

published in the last decades (Wang et al., 2010; Liu et al., 2011; Liu et al., 2014;

Lorente, et al., 2016). In addition to plant growth and development, light quality

also affects fruit ripening. For this reason, much attention has been paid to improve

the nutritional quality of fruit such as tomato using the genetic and molecular apprby

manipulating light signal transduction machinery (Liu et al., 2014; Lorente, et al.,

2016).

Tomato is one of the most popular vegetables in the world, which can be grown

in the field, but increasingly, greenhouse tomato production occurs very commonly.

In some countries, tomatoes are almost exclusively grown in greenhouse (Brazaitytè

et al., 2009). With rapid advances in lighting technology, tomato can be grown all

year around in greenhouse, thus, allowing for stable tomato production to meet its

highly market demand, regardless of the environmental conditions. Among the var-

ious artificial light sources, light-emitting diodes (LEDs) is the most promising

supplemental lighting technology for greenhouse crop production due to its small

size, long lifetime, narrow bandwidths and cool emitting temperature (Muneer et

al., 2014).

Plants require light for photosynthesis as the primary energy source. Changes

in light intensity and quality directly influence the ability of phototrophs to fix CO2

and thus require photosynthetic cells to adjust their cellular metabolism accordingly.

As sessile organisms, plants have tremendous capacity to adjust their metabolism in

response to changes in growth conditions. To obtain a holistic picture of the light-

driven metabolism, it is highly needed to develop a systematic approach to under-

stand the metabolic changes and adjustments of plants. In this regard, constraint-

based modelling (CBM) has proven to be a very promising tool for the systems-

level analysis of cellular metabolism, yielding new insights into condition-depend-

ent metabolic flux distributions and cellular physiology. As yet, such constraint-

based metabolic models have been constructed for a variety of biological systems,

including photosynthetic organisms. Moreover, to characterize metabolic properties

underlying changing light conditions, attempts have already been made (Chang et

Chapter 5

80

al., 2011; Vu et al., 2012; Poolman et al., 2013; Lakshmanan et al., 2015). For ex-

ample, Poolman et al. (2013) studied the impact of light intensity on rice metabolism

using a genome-scale metabolic model, in which the different roles of mitochondrial

metabolism in a leaf were addressed. To model the metabolic influence of different

light spectral ranges on photosynthetic metabolism, Chang et al. (2011) developed

a novel approach that allows a stoichiometric description of a given lighting source.

These so-called ‘prism reactions’ enable decomposition of different light sources to

the optimal absorption wavelength ranges of the light utilizing reactions, thereby

providing valuable insights into light-driven metabolism. Very recently, this ap-

proach was employed to the work of rice, accounting for the metabolic diversity

under different light conditions (Lakshmanan et al., 2015).

All these studies have certainly enhanced our understanding of metabolic ad-

justments and adaptions to changing environmental conditions, however, most

researchers often study the effect of light intensity or light quality independently,

and very few studies have systematically examined the influence of changing light

intensity on plants grown under different light sources. Moreover, there is a lack of

information on the regulation of cellular energy (ATP) and the reducing equivalent

NADPH productions, i.e. ATP/NADPH ratio. ATP/NADPH ratio represents a key

fine-tuning parameter in light energy conversion, their production and consumption

must be balanced to improve photosynthetic productivity and prevent photoinhibi-

tion or photodamage (Kramer and Evans, 2011; Walker et al., 2014). Plants possess

multiple photosynthetic electron transport flows (PETs) to drive the production of

ATP and NADPH, of which the photosynthetic linear electron flow (LEF), produc-

ing ATP and NADPH in a ratio of 1.28 (9 ATP and 7 NADPH), is the most

extensively used pathway (Allen, 2002). However, Calvin cycle driven by photo-

phosphorylation requires an ATP/NADPH ratio of 1.5 (1 molecule of CO2 consumes

3 ATP and 2 NADPH), additional ATP is hence required. To this end, plants have

developed a number of alternate electron flow (AEF) pathways, including cyclic

electron flow (CEF), Mehler’s reaction and chlororespiratory NAD(P)H dehydro-

genases reaction (NDH), to recycle electrons resulting in a net production of ATP

without generation of NADPH (Nogales et al., 2012). CEF has been proposed to

balance the ATP/NADPH ratio working alongside LEF (Allen, 2003), whereas the

precise role of AEF pathways and the extent to which they are used and cooperate

amongst themselves are still not clear. In this chapter, it is therefore very interesting

to study how changes in light intensity and quality influence cellular growth and

how the ATP/NADPH ratio is regulated accordingly. To this end, a light-specific

metabolic model of tomato for seven different light sources was constructed, and

Modelling the effects of light intensity and quality

81

the metabolic states under low light (LL, 50 µmol/m2/s) and high light (HL, 700

µmol/m2/s) were predicted for each light source, respectively.

5.2 Methods

5.2.1 Reconstruction of a tomato light-specific model

The tomato light-specific model was reconstructed by expanding our model

iHY3410 described in Chapter 3 to describe the photon-utilising reactions and prism

reactions in a wavelength-specific manner as implemented by Chang et al. (2011)

and Lakshmanan et al. (2015), respectively. In the tomato light-specific model, the

photosynthetic active light spectrum (PAR, 400-700 nm) was divided into 30 parts,

each denoting 10 nm cumulative region. For instance, the first part denoted by 405

nm covers the region from 400 to 410 nm, the second part denoted by 415 nm covers

from 410 to 420 nm, and so on. The prism reaction, which represents the wave-

length-specific photon composition of a given light source, was generated according

to the approach developed by Chang et al. (2011). Mathematically, it can be repre-

sented as below:

where C is the ratio of photon available in that particular range to total photons in

the visible spectrum of a given light source (Chang et al., 2011). In this manner, it

is possible to distinguish the amount of emitted wavelength-specific photons that

drive different metabolic reactions.

Prism reactions for 7 distinct light sources were generated, including solar

light, white light, blue, green, yellow, and red peak at 640nm (Red640) as well as

red peak at 660nm (Red660) LEDs using the manufacturer’s spectral irradiance

graphs. The spectral distribution of white light, blue, green, yellow and Red640 were

obtained from Liu et al. (2011), while the spectrum of Red660 was collected from

Yang et al. (2014). For solar light, its spectral irradiance was collected from an

online resource (http://www.mv.helsinki.fi/aphalo/photobio/lamps.html). The re-

sulting prism reaction equations were then added to iHY3410 to enable light source-

specific simulations.

PhotonVis 410 420 430 440 450

400 410 420 430 440

460 470 480 490 500

450 460 470 480 490

510

500

C photon405 C photon415 C photon425 C photon435 C photon445

C photon455 C photon465 C photon475 C photon485 C photon495

C photon505

520 530 540 550

510 520 530 540

560 570 580 590 600

550 560 570 580 590

610 620

600 610 6

C photon515 C photon525 C photon535 C photon545

C photon555 C photon565 C photon575 C photon585 C photon595

C photon605 C photon615 C

630 640 650

20 630 640

660 670 680 690 700

650 660 670 680 690

photon625 C photon635 C photon545

C photon655 C photon665 C photon675 C photon685 C photon695

Chapter 5

82

Accordingly, the light-utilising reactions, including photosystems I (PSI), pho-

tosystems II (PSII) and protochlorophyllide oxidoreductase were drafted, in a

wavelength-specific manner. For example, PSII at 405 nm is written as follows:

4.0 photon405_Plas + 2.0 PLASTOQUINONE_Plas --> OXYGEN-

MOLECULE_Plas + 2.0 PLASTOQUINOL-1_Plas + 4.0 Pumped-PROTON_Plas

The available wavelength-specific photons for each light-utilising reaction was

constrained by the maximal absolute amount of usable photons which is derived

based on the absorptivity graphs (Shioi and Sasa, 1984; Nield et al., 2000; Kargul

et al., 2003) multiplied with the total actual photons (320 µmol/m2/s; Liu et al.,

2011). More information is given in Appendix 5.6.1. Excess wavelength-specific

photons which are not metabolically utilised by light-utilising reactions leave the

network via so-called sink reactions, for example, ‘DM_Photon405’.

5.2.2 Constraint-based flux analysis

All simulations in this chapter were performed using ScrumPy package for Linux.

FBA and FVA were performed as described in Chapter 3 (section 3.2.3). A two-step

optimization procedure was applied, where the first step is to maximize growth and

the second step is to minimize total flux. The computation is performed by adding

the objective value, i.e. growth rate of the first optimization as an additional con-

straint to the second optimization. The model was constrained with values for

sucrose and amino acid export to the phloem, chlorophyll turnover (Appendix

5.6.1), nitrate uptake, ATP maintenance cost, a 3:1 ratio for the carboxylase and

oxygenase activities of Rubisco, and other C3-specific constraints as described in

Chapter 3. Due to the scarcity of the experimental data for different light sources,

the phloem composition was assumed remain constant for each light source, thus,

allowing to assess the metabolic changes under different light sources in order to

produce the required metabolites. It should be noted that chlorophylls are also taken

into account in the output demand, although they are not stored in the phloem sap.

This is because chlorophyll can be damaged at high light intensities, resulting in

photoinhinition by decreasing the photosynthetic capacity. Thus, we assumed there

is a constant need to synthesize chlorphyll, i.e. chlorophyll turnover. Chlorophyll

turnover rate were obtained from a previous publication (Nunes-Nesi et al., 2007),

and assumed to be constant under each light source. In addition, fluxes through the

chloroplastic NADPH dehydrogenase (NDH) and plastoquinol oxidase (PTOX)

were set to zero as chlororespiration, involved in NDH and PTOX, is thought to

Modelling the effects of light intensity and quality

83

occur primarily in the dark (Josse et al., 2000; Yamamoto et al., 2011). As men-

tioned above, a set of constraints for the absorption of light-utilising reactions, at

specific wavelength were also defined (detailed in Appendix 5.6.1).

To relate the model predictions to physiological values, the light saturation

point (LSP) of each light source was scaled to a respective experimentally measured

LSP for tomato plants grown under approximate 320 µmol/m2/s (Liu et al., 2011).

5.2.3 Flux-sum

At steady-state, the net rate for internal metabolite consumption and production is

zero, but its turnover rate is not. To compare the metabolite turnover rates among

different light conditions, the ‘flux-sum’ of metabolite i, is used as a proxy for

metabolite turnover. In mathematical, the flux-sum can be formulated as:

where Sij is the stoichiometric coefficient of metabolite i involved in reaction j, vj

denotes metabolic flux of reaction j (Chung and Lee, 2009).

5.3 Results

5.3.1 Representation of photosynthetic light reactions in existing

plant GSMs

As mentioned earlier, plants possess multiple PETs such as LEF and AEF to main-

tain the variability of ATP/NADPH ratio. However, representing these

photosynthetic light reactions as biochemical reactions is a big challenge for con-

structing the metabolic network of photosynthetic organisms because of the

complicity of light capture. This would lead to the representation of light reactions

nonstandard and non-conserved. As shown in Table 5.1, the reconstructed plant

metabolic networks do differ in the representation of light reactions, in particular,

on the degree of details in representing light reactions. Among the 22 surveyed plant

metabolic networks, only 2 models accounted for all the possible PETs and detailed

i

0.5 | |i ij j

j

S v

Chapter 5

84

Table 5.1 Electron flow and reactions taking place during the light phase of photosynthesis in the existing large-scale metabolic models of plants.

Species No. Model LEF

ATPase AEF

NGAM PSII Cytb6f PSI FNR FQR NDH PTOX Mehler PHOTOR

Arabidopsis

1 Poolman et al., 2009[1] - - - - - - - - - - √

2 de Oliveira Dal’Molin et al., 2010 Merged LEF & CEF - - √ - - - - - √ -

3 Radrich et al., 2010[2] - - - - - - - - - - -

4 Saha et al., 2011 Merged LEF & CEF - - √ - - - - - √ -

5 Mintz-Oron et al., 2012 Merged LEF & CEF - - √ - - - - - √ -

6 Chung et al., 2013 Merged LEF & CEF - - √ - - - - - √ -

7 Cheung et al., 2013 √ √ √ √ √ √ √ √ √ √ √

8 Arnold and Nikoloski, 2014 √ √ √ √ √ - - - - √ -

Barley

9 Grafahrend-Belau et al., 2009[3] - - - - - - - - - - -

10 Grafahrend-Belau et al., 2013[4] Lumped LEF - - √ √ Lumped CEF - - - √ -

Rapeseed

11 Hay and Schwender, 2011 √ √ √ √ √ √ - - - √ √

12 Pilalis et al., 2011 Merged LEF & CEF - - - - - - - - - -

Maize

13 de Oliveira Dal’Molin et al., 2010b Merged LEF & CEF - - √ - - √ - - √ -

14 Saha et al., 2011 Merged LEF & CEF - - √ √ - - - - √ -

15 Simons et al., 2014 √ √ √ √ √ - - - - √ -

Sorghum 16 de Oliveira Dal’Molin et al.,2010b Merged LEF & CEF - - √ - - √ - - √ -

Sugarcane 17 de Oliveira Dal’Molin et al., 2010b Merged LEF & CEF - - √ - - √ - - √ -

Rice

18 Poolman et al., 2013 Lumped LEF - - - - Lumped CEF - - - √ √

19 Lakshmanan et al., 2013 Merged LEF & CEF - - - - - - - - √ -

20 Lakshmanan et al., 2015[5] √ √ √ √ √ √ √ √ √ √ -

Tomato 21 Colombié et al., 2014[6] - - - - - - - - - - √

22 Yuan et al., 2016 √ √ √ √ √ √ - - - √ √

Modelling the effects of light intensity and quality

85

Symbol “-” indicates information is not included; “√”indicates information is included. LEF, linear electron flow;

AEF, alternative electron flow; PSII, photosystem II; Cytob6f, cytochrome b6f complex; PSI, photosystem I; FNR,

ferredoxin NADP+ oxidoreductase; ATPase, plastidic ATP synthethase; FQR, ferredoxin PQ reductase; NDH,

NAD(P)H dehydrogenase; PTOX, plastid terminal oxidase; Mehler, Mehler’s reaction; PHOTOR, photorespira-

tion. [1] the model represents a heterotrophic cell; [2] the work aimed at presenting a methodology for integration of dif-

ferent databases, rather than FBA simulation; [3] the model was used for studying the seed metabolism; [4] the study

constructed an tissue-specific models (including barley leaf, stem and seed) on a whole-plant scale. The indicated

light-dependent reactions are present in the leaf and stem models; [5] the model is an extension of a core model

published by Lakshmanan et al. (2013); [6] the model was used to investigate the heterotrophic metabolism of tomato

fruit.

them into several reactions (Cheung et al., 2013; Lakshmanan et al., 2015). Notice-

ably, 10 models simply used an aggregate reaction to describe

photophosphorylation, in which photons convert H2O into ATP and NADPH with a

fixed ratio of 1.5. Although such representation accounts for both LEF and AEF,

however, it is not sufficient to infer which AEF exactly contributes to the additional

ATP production in these studies. In some other studies, light reactions are detailed

into several reactions, but only accounting for LEF and CEF (Arnold and Nikoloski,

2014; Yuan et al., 2016).

Undoubtedly, an accurate representation of photophosphorylation will promote

to better understand photosynthetic processes. Therefore, it necessitates to take into

account all possible PETs. To this end, the tomato genome-scale model, iHY3410

described in Chapter 3 was extended by adding the ferredoxin plastoquinone reduc-

tase (FQR) reaction as well as the Mehler and related reactions, e.g., NDH and

PTOX as putative additional alternate electron transport pathways (Appendix 5.6.2).

5.3.2 Two published light-specific models

To build a light-specific model of tomato, two references from published models

were exploited: one is for Chlamydomonas (iRC1080; Chang et al., 2011), and the

other is for rice (iOS2164; Lakshmanan et al., 2015). Both of the models accounted

for the light-driven photophosphorylation reaction in a wavelength-specific way,

and permitted growth simulations under different light sources. The modelling ap-

proaches used in algae and rice (illustrated in Figure 5.1) are quite similar, but with

a few differences that are discussed in detail below.

Chapter 5

86

Figure 5.1 Comparison of light-specific modelling approaches between algal model and rice model.

In addition to the traditional reactions (e.g., transporters and metabolic reac-

tions) included in metabolic networks, 3 types of reactions, i.e. prism reaction,

photon-utilising reaction and sink reaction, were introduced to account for the reac-

tions that associate with light in the light-specific models. Prism reaction described

in a stoichiometric way represents the spectral composition of a given lighting

source. Either in Chang et al. (2011) or Lakshmanan et al. (2015), the calculation

of the coefficient for each photon wavelength bandwidth in prism reaction (denoted

as C) is the same, defined as the number of photons available at a particular wave-

length bandwidth divided by the total available photons in PAR emitted by a given

light source, however, the two models differed in decomposing the lighting spec-

trum. In Chang et al. (2011) approach, the spectrum of light source is decomposed

by the most efficient spectral bandwidths (blue and red regions) based on the reac-

tion activity of photon-utilising reactions in an unequal and non-continues way. It

should be noted that this binning manner results in overlapping regions, hence, light

sources are distinguished by the amount of photons that are utilised by metabolic

reactions. In contrast, the rice model iOS2164 divided the lighting spectrum into

equal parts through the whole visible light range in a continuous way. Thus, the sum

of C (i.e. 1) remain constant among all the light sources, indicating that light sources

are distinguished by the amounts of wavelength-specific photons rather than by the

amount of photons.

Photon-utilising reaction represents the reactions that are dependent on light

such as photosystem I (PSI) and photosystem II (PSII). Given that not all absorbed

photons are utilised by metabolic reactions and these photons must be leaked from

the system to avoid photodamage to the cells, sink reaction denoting the absorbed

Modelling the effects of light intensity and quality

87

but not utilised photons is added to the light-specific model. Additionally, light ab-

sorption reaction is added to the model of Lakshmanan et al. (2015), in a

wavelength-specific manner as described the prism reaction. Light absorption reac-

tion is the intermediate step between light input and the respective photon-utilising

metabolic reactions in the rice model, which is combined into photon-utilising re-

action in the alga model. One thing that needs to be pointed out is that in rice model,

wavelength-specific photons that are not absorbed by one light absorption reaction

will directly leave the system via sink reaction, which is actually not realistic in

planta because photons that are not absorbed by one light absorption reaction could

be absorbed by other photon absorption reactions. Consequently, changes to the ap-

proach introduced in rice model were made, by combing the light absorption

reaction with light-utilising reaction as implemented by Chang et al. (2011), mean-

while, light-utilising reac-tions are constrained by the maximal usable photons, at

specific wavelength (section 5.2.1).

5.3.3 Model properties

The genome-scale metabolic model of tomato was extended, iHY3410 described in

Chapter 3, to a light-specific model. In total, the developed model includes 127 re-

actions denoting the wavelength-specific spectral decomposition (for detailed

procedure see section 5.2.1). As the photo-synthetic reactions are of particular im-

portance, all photosynthetic reactions in the iHY3410 were corrected where

necessary. Specifically, the protochllidelide reductase reactions were corrected,

changing from the cytosol to the plastid. The resulting tomato light-specific model

is given in Appendix 5.6.3, and a simplified scheme of the tomato light-specific

model is shown in Figure 5.2.

Chapter 5

88

Figure 5.2 Schematic overview of the light-specific model of tomato mature leaves. The numbers (15 and 2) indi-

cate the number of associated reactions.

5.3.4 Growth simulation of tomato cells under different light

conditions

On the basis of the reconstructed light-specific model, the metabolism of mature

non-growing leaves was studied by maximizing growth rate as a function of light

influx. Optimal growth rates were computed under changing light intensity for a

given light source while subject to constraints mentioned above (section 5.2.2). Ba-

sically, constraining the upper limit on wavelength-specific photons is quite similar

to setting an upper limit on CO2 uptake rate as implemented in previous studies

(Nogales et al., 2012; Maarleveld et al., 2014). While in this case, both maximal

light absorption and utilisation are taken into account in model constraints.

According to the LSP, two unique states are identified for each given light

source: (1) low light (LL) in which light is limiting for growth, and (2) high light

(HL) in which light is excess, but CO2 is limiting for growth (correspond to carbon-

limiting sate as carbon fixation becomes light-saturated). To enable a fair compari-

son among different light sources, a light intensity of 50 µmol/m2/s, which is lower

than the LSPs of all considered light sources to represent LL state, was selected for

Modelling the effects of light intensity and quality

89

further exploration. Likewise, a light intensity of 700 µmol/m2/s is defined as HL

state. A detailed overview of the characteristics of these scenarios is given in Ap-

pendix 5.6.4.

As expected, growth rate increased with the light intensity, which holds true

for all considered light sources (Table 5.2). Under LL state, the growth rate was

ranked as follows: white > solar > blue > Red660 > Red640 > green > yellow. Sim-

ilarly, solar and white had the highest growth rate and yellow had the lowest under

HL state. Unexpectedly, under HL, the growth rate of green is even higher than

Red640, but lower than Red660.

Table 5.2 A summary of simulation results predicted from all studied conditions. фCO2, moles of CO2 fixed per

mole of absorbed photons; Vi/Vo, photon conversion efficiency; LL, low light; HL, high light.

Simulated

Conditions

Growth Rate

(µmol/m2/s) фCO2 Vi/Vo (%)

Light Utilisation Efficiency

µmol/photon µmol/J

Solar LL 0.090 0.060 99.0 0.0018 0.0086

HL 0.270 0.077 16.4 0.0023 0.0098

White LL 0.105 0.070 99.1 0.0021 0.0090

HL 0.270 0.077 16.4 0.0023 0.0098

Blue LL 0.074 0.065 74.5 0.0020 0.0077

HL 0.129 0.072 8.4 0.0022 0.0084

Green LL 0.022 0.044 33.6 0.0013 0.0057

HL 0.058 0.062 4.4 0.0019 0.0079

Yellow LL 0.010 0.029 24.2 0.0008 0.0042

HL 0.021 0.043 2.4 0.0013 0.0063

Red640 LL 0.034 0.053 43.4 0.0016 0.0084

HL 0.050 0.060 4.0 0.0018 0.0096

Red660 LL 0.058 0.062 62.3 0.0019 0.0101

HL 0.069 0.065 5.1 0.0019 0.0105

It should be noted that in this study, all the photons emitted from a light source

are assumed to be incident upon a cell because there is little quantitative data avail-

able on the amount of light that is incident upon a cell. Due to the occurrence of

non-productive dissipation of absorbed photons, e.g., dissipated as heat or re-emit-

ted as chlorophyll fluorescence, these incoming photons are not completely

metabolicallyabsorbed. In this study, the photon conversion efficiency (Vi/Vo) ac-

counting for the amount of photons that are metabolically utilised varied with the

studied conditions. Unsurprisingly, the photon conversion efficiency is significantly

decreased at HL, the light-saturated condition. For yellow light source, only 2.4%

of the incident photons are used for the metabolic process at HL state, while the

majority are leaked from the system.

Chapter 5

90

Going beyond merely comparing the growth rate under different light condi-

tions, the light utilisation efficiency of these light sources was evaluated in terms of

three criteria, including quantum field (фCO2), and growth rate on numbers of pho-

tons as well as on photon energy. The quantum yield of photosynthesis is regarded

as a definitive and also generic measure of the energetic efficiency of phototrophs.

In this study, the quantum yield is defined by the moles of CO2 fixed (by Rubisco

carboxylase) per mole of absorbed photons (Appendix 5.6.5). The computed quan-

tum yields are in a range of 0.029 to 0.077 in the studied conditions, within which

the larger value is obtained from white, followed by solar and blue, and yellow is

the lowest, irrespective of LL or HL. In term of the growth rate on the number of

photons, similar results are obtained for the studied conditions. However, from the

perspective of energy, it reveals a different story, showing the two red lights (espe-

cially Red660) are more efficient compared to other light sources no matter the LL

or HL state.

5.3.5 Exploring energy balancing in the light

Analyses of light utilisation efficiency gave us additional information on the cellular

metabolism, e.g., flux distributions between PSII and PSI. The same amount of

fluxes were predicted to flow into PSI and PSII (i.e. PSI/PSII = 1) in all the simu-

lated conditions (Appendix 5.6.5), suggesting ATP is synthesized in the plastid

entirely through LEF, i.e. CEF around PSI is inactive. As aforementioned, LEF

alone is unable to satisfy the ATP/NADPH ratio requirement. Hence, there should

be other routes that contribute to the balance of ATP and NADPH in the model

predictions. Both FBA and FVA results showed no AEF is active in each simulated

condition. However, knocking out mitochondrial complexes would substantially de-

crease the growth rate, take LL state of white light as an example (Table 5.3),

revealing mitochondrial respiration is directly involved in optimizing growth per-

formance in the light. Further analyses showed that the energetic interaction

between plastid and mitochondria proceed by three shuttles that export excess re-

ducing equivalents from the chloroplast to the cytosol, then balance the

ATP/NADPH ratio in the plastid, rather than by the mitochondrial ADP/ATP trans-

locator that exports ATP out of the mitochondria in exchange for ADP.

Modelling the effects of light intensity and quality

91

Table 5.3 Effects of mitochondrial ETC knockouts (KO) on fluxes through the compared reactions (in case of white

at LL state).

ATPase_Mito, mitochondrial ATP synthase, i.e. Complex V; ATPase_Plas, plastidic ATP synthase; symbol “-”

indicates no flux. Predicted fluxes are in µmol/m2/s.

Malate-Oxaloacetate shuttle

The malate-oxaloacetate (Mal-OAA) shuttle is a malate valve, comprised of a mal-

ate/OAA transporter and some isozymes of malate dehydrogenase (MDH). In the

Mal-OAA shuttle, excess plastidic NADPH reduces OAA to malate, which is then

transported to the mitochondria and further used in the mitochondrial electron

transport chains (ETCs). Our model predicted that NADP+-dependent MDH con-

verts OAA to malate in the plastid, whereas an NAD+-MDH operates in the

mitochondria, generating NADH which is directly oxidized via the ETCs (Figure

5.3). This shuttle was found in all studied conditions.

Figure 5.3 Mal-OAA exchange among plastid in all scenarios. OAA, oxaloacetate; MDH, malate dehydrogenase;

Mal, malate; ETC, electron transport chain.

Exchange of carbon compound among plastid

Growth Rate ATPase_Mito ATPase_Plas AEF

Wild type 0.105 1.362 5.305 -

KO Complex I 0.099 1.119 5.305 -

KO Complex III 0.094 0.917 5.305 -

KO Complex IV 0.094 0.917 5.305 -

KO ATP/ADP translocator 0.105 1.362 5.305 -

Chapter 5

92

In addition to the Mal-OAA shuttle, our model simulations showed that, at HL states

of solar and white lights, chloroplasts can also interconnect with other organells by

exchanging carbon compounds (Figure 5.4). During this exchange, the malate-cit-

rate shuttle exports carbon from the peroxisome. The 2-oxoglutarate (2-OG)

produced by cytosolic citrate dehydrogenase serves as a substrate for synthesis of

glutamate in the chloroplast. The FVA results confirmed that this route is always

active in the HL state of solar and white, but is optional for other conditions.

Mitochondrial aldehyde dehydrogenase

In the LL state of yellow light, the model predicted that the aldehyde dehydrogenase

(LADH) involved in the energetic coupling between the plastid and the mitochon-

dria (Figure 5.5). LADH metabolizes acetaldehyde produced from threonine into

acetate. The NADH produced by mitochondrial ALDH goes to the mitochondrial

ETC.

Figure 5.4 Carbon compound exchange among plastid in HL of solar and white. OAA, oxaloacetate; MDH, malate

dehydrogenase; Mal, malate; ETC, electron transport chain; 2-OG, 2-oxoglutarate; Glu, glutamate; Fdred, ferredoxin

reduced; Fdox, ferredoxin oxidized; GOGAT, glutamate synthase.

Figure 5.5 Function of ALDH under LL state of yellow. ETC, electron transport chain; CS, citrate synthase; Ac-

CoA, acetyl CoA; ACET, acetate; ACETALD, acetaldehyde; ALDH, aldehyde dehydrogenase.

Modelling the effects of light intensity and quality

93

5.3.6 Flux-sum differences in cofactors ATP, NADH and NADPH

The flux-sum for a metabolite (see section 5.2.3) can be considered as a proxy for

the metabolic pool size at steady-state (Chung and Lee, 2009). This approach has

been used to elucidate the metabolic state of plant systems (Simons et al., 2014;

Basler et al., 2016). In this chapter, the flux-sum range of compartmentalized spe-

cific cofactors ATP, NADPH and NADPH were computed from the resulting flux

distributions at the optimal solution under different light conditions (Appendix

5.6.5). We observed that the flux-sum of the mitochondrial and peroxisomal

NADPH were observed to be 0 at the optimal solutions, which implies that NADPH

is not consumed nor produced by any of respective reactions in the mitochondria

and peroxisome. For the remainder of selected metabolites, the highest level for the

flux-sum was obtained from HL state of solar and white light, nonetheless, the low-

est was from LL state of yellow light.

The flux-sum levels for the compartmentalized-specific cofactors were sub-

stantially higher in the HL scenarios for each light source. However, for all the

considered light sources, the flux-sum levels of mitochondrial NADH and mito-

chondrial ATP were relatively higher in the LL states compared to those of cells in

HL states, reflecting the contributions of mitochondria to energy balance is bigger

under LL state (Table 5.4).

Table 5.4 The flux-sum ranges of mitochondrial cofactors. lb, lower bound; ub, upper bound.

Light Source Cofactor LL HL

lb ub lb ub

Solar NADH_Mito 2.24 2.25 1.74 1.74

ATP_Mito 5.32 5.32 3.34 3.34

White NADH_Mito 1.68 1.70 1.74 1.74

ATP_Mito 4.09 4.09 3.34 3.34

Blue NADH_Mito 7.05 7.06 6.68 6.71

ATP_Mito 16.44 16.44 15.84 15.84

Green

NADH_Mito 0.48 0.48 0.44 0.44

ATP_Mito 1.15 1.15 1.08 1.08

Yellow NADH_Mito 1.95 1.95 1.92 1.92

ATP_Mito 4.51 4.51 4.46 4.46

Red640 NADH_Mito 1.88 1.88 1.83 1.84

ATP_Mito 4.40 4.40 4.33 4.33

Red660

NADH_Mito 1.81 1.82 1.78 1.79

ATP_Mito 4.30 4.30 4.25 4.25

Chapter 5

94

5.4 Discussion

In this chapter, we aim to in silico study the effects of light intensity and light quality

on the cellular growth and metabolism of tomato leaves. To this end, a light-specific

model for tomato was built by describing the photon-utilising reactions in a wave-

length-specific way. Similarly, the given light sources (in total, seven) were

metabolically represented in a wavelength-specific manner.

5.4.1 Influence of varying light on the cellular growth

The model predicted that the growth rate at HL states for each given light source is

higher than its respective value at LL states (Table 5.2). Irrespective of the changes

in light intensity (LL or HL), the higher growth rate was obtained from solar, white

and blue lights, while the lowest value was found from yellow light. These predic-

tions were supported by previous findings that the growth of tomato plants irradiated

with monochromatic lights, in particular, yellow and green lights, was inhibited as

compared with that of plants grown under full spectrum lights such as white (Wu et

al., 2014; O’Carrigan et al., 2014; Liu et al., 2014). For example, Wu et al. (2014)

found that tomato plants under white light showed higher growth rate compared

with other monochromatic light treatments including blue, green, yellow and red,

whereas, the growth of tomato plants under green and yellow light treatments were

severely inhibited. Liu et al. (2014) reported that tomato plants with yellow light

treatment had lowest fresh weight and dry weight. These effects of monochromatic

lights on tomato plants have also been observed in other plants (Yu and Ong, 2003;

Wang et al., 2009). For instance, the results of Wang et al. (2009) showed that all

cucumber plants grown under monochromatic lights had reduced growth and CO2

assimilation rate as compared with plants grown under white light.

It is notable that the light spectra in many reported experiments, which were

produced either by fluorescents or LEDs, were often inconsistent, thus often leading

to mixed results. For example, the results of Wu et al. (2014) showed that the growth

of tomato plants had no significant differences between white and red lights (with a

maximum intensity at 660 nm, i.e. Red660), although plants treated with white light

had relatively higher growth than those of plants irritated with red light; however,

in the work of Liu et al. (2014), the growth reduction of tomato plants under red

light (with a peak at 640 nm, whereby Red640) was more significant in comparison

with those of white light. Previous evidence has proved that early LEDs were red

with the most efficient emitting at 660 nm, which is close to an absorption peak of

chlorophyll (Sager and McFarlane, 1997). Not unexpectedly, our model predicted

Modelling the effects of light intensity and quality

95

that the cellular growth rate of Red640 differed from that of Red660 (Table 5.2).

Given that the influence of monochromatic lights on photosynthetic capacity was

relatively limited compared with full spectrum light such as solar and white, at-

tempts have been made toward investigating the response of plants to the

combination of different light qualities such as red, blue and green lights. And nu-

merous studies have proven that the combination of red and blue LEDs is an

effective lighting source for several plants (Yorio et al., 2001; Kim et al., 2005; Fan

et al., 2013).

As mentioned earlier, the quantum yield is a measure of light utilization effi-

ciency. A theoretical maximal value of фCO2 is 0.111 for C3 plants, while, the average

actual value was reported to be 0.052 under ambient conditions (Skillman, 2008).

Variation in light environments such as light intensity and light quality is expected

to affect the quantum yield. Unsurprisingly, some computed фCO2 (e.g., of green and

yellow) was lower than the value of 0.052. Our prediction that white light had a

higher quantum yield is in line with other studies (Zhang et al., 1997). Despite solar

and white lights had higher growth rate relative to monochromatic light treatments,

while the latter, e.g., red lights are more efficient for converting electricity to pho-

tosynthetically fixed carbon because the conversion of electrical power to photons

increased with wavelength.

Apart from the growth, it is known that the monochromatic lights had signifi-

cant influences on morphogenesis of plants (Wang et al., 2009; Wu et al., 2014).

However, the structure and physiology of plants are regulated, in part, by light sig-

nals, which is not considered in this present study.

5.4.2 Energy balancing under varying light conditions

Chloroplasts and mitochondria play important roles in meeting the energy demands

and maintaining cellular redox balance (Griffin and Turnbull, 2012). In an illumi-

nated plant cells, metabolic processes occurred in these two organelles are

interdependent but also mutually beneficial. The metabolic interactions between

these two organelles have been reported in several studies (Bailleul et al., 2015;

Cheung et al., 2015; Kim et al., 2016; Levering et al., 2016). Consistent with these

reported findings, our model predictions also support the essential role of mitochon-

drial metabolism in dissipating excess redox equivalents from chloroplast, thus

regulating the ATP/NADPH ratio for sustaining photosynthetic carbon assimilation

(Table 5.3). Additionally, the performed flux-sum analysis of cofactors revealed that

the contribution of mitochondrial respiration is bigger at LL state compared to that

of HL state (Table 5.4).

Chapter 5

96

The possible routes for interaction between chloroplasts and mitochondria have

been proposed (Hoefnagel et al., 1998; Blanco et al., 2014). In this study, the mech-

anisms that are used to balance the ATP/NADPH ratio during the communication

were extensively explored. As reported in the other studies (Bailleul et al., 2015;

Levering et al., 2016), the Mal-OAA shuttle was predicted to be involved in redox

shuttling between the chloroplast, mitochondria, and cytosol, in all the modelled

conditions (Figure 5.3). Nonetheless, the ATP produced by mitochondrial ETC was

not transported back to chloroplast via the plastidic ADP/ATP translocator in our

predicted results (Table 5.3).

In addition to the Mal-OAA shuttle, our model predictions indicated that the

balancing of ATP/NADPH ratio can also be accomplished through the exchange of

carbon compound among the plastid at HL states for solar and white lights (Figure

5.4), and the mitochondrial ALDH at LL state of yellow (Figure 5.5). ALDH was

reported to play a beneficial role when plants suffer abiotic stress, in particular,

drought and flood (Nakazono et al., 2000; Brocker et al., 2013). Generally, lower

light is not a well-studied abiotic stress compared with other stresses such as drought

because its effect is not fatal to plants. ALDH reaction produces NADH during the

conversion of acetaldehyde to acetate. Considering that plants under yellow light

treatment often had lower growth and photosynthetic capacity, it is likely that the

enzyme is active when cells require regeneration of reducing power.

In summary, the discussions above demonstrate that plant cells possess multi-

ple mechanisms to cope with fluctuations in the light environment, implying the

metabolic flexibility of plant cells in response to different growth conditions.

5.4.3 Potential limitations

In constraint-based metabolic models, maintenance energy is divided into two major

components: the growth-associated maintenance (GAM) and non-growth-associ-

ated maintenance (NGAM). It is relatively easy to estimate GAM, but the

experimental determination of NGAM is more difficult (Sweetlove et al., 2013),

thus resulting in a large variation in NGAM within the published plant flux-balance

metabolic models (Table 5.1).

The accuracy of FBA predictions highly depends on the model structure and

constraints, of which NGAM has the strongest effect on model prediction compared

with the remaining constraints (Cheung et al., 2013). NGAM accounts for energetic

cost for biosynthesis-related processes, including metabolite transport, protein and

nucleic acids polymerisation and processing and protein translocation. In many or-

ganisms, a significant amount of cellular ATP is expected to be used to sustain

Modelling the effects of light intensity and quality

97

cellular functions that are not directly associated with biomass synthesis and growth.

Hay and Schwender (2011) reported that NGAM consumes up to about 50% of the

total ATP production. Masakapalli et al. (2010) estimated this fraction to be larger

than 85% for Arabidopsis cell cultures. In a study of tomato fruit, the ATP con-

sumption used for maintenance was reported to be in a range of 22-29%, regardless

of the development stage (Colombié et al., 2015). Given the substantial effect of

NGAM on predicted flux distributions (Feist et al., 2007; Poolman et al., 2009;

Cheung et al., 2013; Kim et al., 2016), it would result in an even larger uncertainty

in model predictions to neglect NGAM as done by many plant metabolic models

(Table 5.1).

In this chapter, ATP maintenance cost, a value of 7.1 mmol g-1 DW h-1 reported

in an Arabidopsis model was exploited (Poolman et al., 2009), and was assumed to

be the same in all the studied conditions because no information is available regard-

ing the ATP maintenance energy for tomato. The proportion of the fixed

maintenance cost accounts for 19-79% of all cellular ATP all the studied conditions.

However, according to the results of Imam et al. (2015), photosynthetic mainte-

nance energy requirements were significantly affected by light intensity. Such an

observation that maintenance cost varies with the environmental conditions is also

reported by Cheung et al. (2013).

In order to investigate how changes in ATP maintenance alter the predicted

flux solutions, as well as flux through PETs and mitochondrial ETCs, the studied

conditions were examined by performing sensitivity analysis of ATP maintenance

implemented as described by Hartman et al. (2014). The model predicted that

changing the ATP maintenance value led to the flux distribution to change. At the

lower ATP maintenance (< 0.60) in LL state of white light, there is no flux through

mitochondrial ETCs, and the fluxes through the mitochondrial ETCs increased with

increasing ATP maintenance (Appendix 5.6.6). Surprisingly, AEFs did not carry

fluxes either at the lower ATP maintenance conditions, implying the energy gener-

ated by LEF in plastid is sufficient for supporting the energy burden. It is worth

noting that only ~1-8% of cellular ATP was used for other purposes than biomass

production under lower ATP maintenance. Obviously, this value is much less that

those reported by other studies (Alonso et al., 2007; Masakapalli et al., 2010; Hay

and Schwender, 2011; Colombié et al., 2015), indicating that the model predictions

obtained from a value of 7.1 mmol g-1 DW h-1 for ATP maintenance are not unreli-

able.

Another limitation of this study is the actual light absorption for metabolic use.

The need to provide an accurate definition of the metabolic light absorption for the

light condition of interest is obvious. However, there is little information on actual

Chapter 5

98

photon absorption flux for metabolic use because such measurements are not prac-

tical. Typically, much of the incident light is reflected or scattered before it is

utilised in photosynthesis, particularly, at high light. In this study, the same upper

limit on wavelength-specific photons were set for all the studied conditions, which

is not realistic actually.

Despite the limitations in this study discussed above, the reconstructed light-

specific model strikingly predicted the different metabolic actives under changing

light conditions using very limited experimental data as constraints. This gives hope

of attaining the goal of addressing both the metabolic response of leaf cells and the

mechanisms regulating ATP/NADPH ratio under various light conditions.

5.5 Conclusions

This work clearly demonstrates the utility of constraint-based modelling to study

the metabolic activities of light-stimulated assimilation in leaf cells and moves our

thinking forward. The present study highlights the importance of mitochondrial res-

piration in optimizing growth in light. In addition, our model predictions suggested

ALDH may represent an important metabolic interaction between plastid and other

intracellular. Nonetheless, it must be stated that the present study should not be con-

sidered the last word in respect of light-driven metabolism investiagtion, but as an

initial step analysing photosynthetic behaviors of plants. Further studies are required

to simulate the effects of different ratios of selected light qualities on the cellular

metabolism. Furthermore, additional information on light quality and phytochemi-

cal production will help us in accurately modelling the effects of light quality on

light-driven metabolism, and also designing the most efficient lightings to promote

plant growth and accumulation of phytochemical concentrations.

Modelling the effects of light intensity and quality

99

5.6 Appendix

Appendix 5.6.1 Model constraints

Table 1: Relative amino-acid composition in phloem sap of tomato leaves. Values obtained from Valle et al. (1998)

are shown as relative molar % of the total amino acids in the phloem, sorted from the most to the less abundant.

Amino-acid molar % of amino-acid in phloem

Glutamine 30.4

Glutamate 15.7

Threonine 8.7

Aspartate 7.6

Phenylalanine 7.2

Serine 5.0

Alanine 4.9

Valine 3.5

GABA[1] 2.9

Lysine 2.8

Leucine 2.6

Isoleucine 2.2

Asparagine 1.9

Glycine 0.9

[1] GABA, γ-aminobutyric acid

Chapter 5

100

Table 2: Constraints for the absorption of light-utilising reactions at specific wavelength (µmol/m2/s).

Wavelength

(nm)

Light-utilising reactions

PSI PSII PCHLDOR DVPCHLDOR

405 3.65408026 3.571196 1.032251 1.032251

415 4.57122079 5.762609 2.709219 2.709219

425 5.09505344 6.411913 3.869828 3.869828

435 6.02854023 7.122096 4.7722 4.7722

445 4.29092208 6.736565 4.290922 4.290922

455 2.91386204 4.463993 2.913862 2.913862

465 2.89691067 4.078462 4.724813 4.724813

475 2.453918 4.200212 2.743555 2.743555

485 2.02729613 3.75381 1.580109 1.580109

495 1.60064973 3.124793 1.105274 1.105274

505 0.94127968 2.029087 0.845721 0.845721

515 0.518035 1.075419 1.274583 1.274583

525 0.38644719 0.689887 1.574335 1.574335

535 0.38589114 0.608729 1.573019 1.573019

545 0.36898065 0.588433 1.743841 1.743841

555 0.41761875 0.547857 2.000655 2.000655

565 0.54817623 0.588433 2.343444 2.343444

575 0.61324236 0.710183 2.255948 2.255948

585 0.71100919 0.852213 2.211398 2.211398

595 0.79242158 0.892797 2.941722 2.941722

605 0.93936621 0.93338 3.671947 3.671947

615 1.0698828 1.13629 4.44521 4.44521

625 1.39654635 1.280821 5.476611 5.476611

635 1.39654635 0.943525 6.852403 6.852403

645 2.00231473 1.623268 8.177219 8.177219

655 3.21445661 1.988503 6.203999 6.203999

665 3.98406376 3.672651 4.136659 4.136659

675 1.64015388 5.072713 2.456764 2.456764

685 0.47608587 3.003051 1.422525 1.422525

695 0.18057753 0.426107 0.947691 0.947691

PSI, photosystem I; PSII, photosystem II; PCHLDOR, chlorophyllide-a:NADP+ oxidoreductase; DVPCHLDOR, divinyl chlorophyllide-a:NADP+ oxi-

doreductase.

Modelling the effects of light intensity and quality

101

Appendix 5.6.2 Schematic representation of photosynthetic machinery

Abbreviations: PSI, photosystem I; PSII, photosystem II; b6f, cytochrome b6f complex; Fd, Ferredoxin; FNR, fer-

redoxin NADP+ reductase; FQR, ferredoxin quinone reductase, NDH, NADPH dehydrogenase; PTOX,

plastoquinol oxidase; Mehler, Mehler’s reaction; PQ, plastoquinone; PC, plastocyanin; Pi, inorganic phosphate. e-

, electrons.

Appendix 5.6.3 Light-specific metabolic model of tomato in ScrumPy and

SBML format. Availabe on the attached files.

Appendix 5.6.4 FBA and FVA results for all the studied conditions. Avail-

abe on the attached files.

Appendix 5.6.5 The calculation of light utilisation efficiency and flux-sum

levels of cofactors. Availabe on the attached files.

Chapter 5

102

Appendix 5.6.6 Responses of mitochondrial ETCs to varying ATP demand

(LL state of white light).

Chapter 6 General discussion

Chapter 6

104

6.1 Summary of contributions

Constraint-based metabolic modelling approach has been proven a powerful tool to

study cellular metabolism at systems-level. With such motivation, this thesis has

initiated a systems modelling approach to understand the metabolic behavior of

plant cells under various growth conditions. The following sections presents the

main contributions of this thesis work.

The examination of how the modelling parameters commonly used in flux-bal-

anced analysis influence the flux distributions in plant metabolic networks (Chapter

2), highlights the significance of model structure when studying the plant metabolic

behavior using flux balance analysis (FBA).

The constructed metabolic model of tomato, iHY3410 (Chapter 3), to our

knowledge, is the first genome-scale metabolic model for simulating tomato leaf

metabolism. The model is capable of successfully describing the metabolic activities

of well-known biochemical pathways for leaf cells under heterotrophic and photo-

trophic conditions.

The application of iHY3410 to investigate the metabolic characteristics under

drought stressed conditions (Chapter 4), presents a starting point for studying the

functional consequences of abiotic stresses. The model predictions that the meta-

bolic flux through the mevalonate (MVA) pathway under drought conditions

significantly differed from that under normal conditions, gave hints to possible

adaptive metabolic responses to drought stresses.

Finally, in Chapter 5, the extension of iHY3410 to a light-specific model allows

us to explore the metabolic changes of tomato leaves in response to changing light

intensity and quality. The comprehensive analyses indicate mitochondrial respira-

tion plays an important role in regulating the balance between ATP and NADPH in

plastid, thus enabling optimal growth performance in varying light conditions. The

interaction between the plastid and other organells is achieved by metabolite shuttles

that differ with the light condition.

6.2 Limitations of this study

Essentially, all models are wrong because they are simplified representations of re-

ality containing limitations and assumptions (Sterman, 2002). Unexceptionally,

constrained-based modelling approach, the central framework of this thesis work,

also has limitations, in particular, when applying it to study phototrophic organisms,

General discussion

105

whose metabolism typically exhibits a diurnal, rather than time-invariant behavior

(Orth et al., 2010; Sweetlove and Ratcliffe, 2011; Ip et al., 2014; Rügen et al., 2015).

Yet, a few attempts have been made to describe temporal variations in phototrophic

metabolism work (Muthuraj et al., 2013; Cheung et al., 2014; Rügen et al., 2015).

In this section, some of the limitations of this thesis work that have not been dis-

cussed earlier are highlighted.

The accurate predictions of flux distributions by FBA highly depends on the

model structure and constraints, amongst which maintenance costs that are not re-

lated to growth is an important parameter. The maintenance costs not only include

the ATP demand but also the reductant expenditure. Including both the ATP and

NADPH maintenance costs has been proved to be important for generating accurate

flux predictions (Cheung et al., 2013). However, this thesis primarily consider the

maintenance costs in terms of ATP consumption, neglecting the NADPH consump-

tion. Moreover, as the values are not known for tomato, the ATP maintenance cost

was set to be the same as that of Arabidopsis in the thesis, irrespective of the study-

ing conditions (see Chapter 5, section 5.4.3).

Additionally, this thesis does not account for secondary metabolites in the bio-

mass constraints. In Chapter 4, our model predicted that isopentenyl pyrophosphate

(IPP), the precursor for the biosynthesis of isoprenoids, is predominantly synthe-

sized via the 2C-methyl-D-erythritol-4-phosphate (MEP) pathway and the

mevalonate (MVA) pathway under the normal and drought stressed conditions, re-

spectively. Isoprenoids, also called terpenoids, are one of the most diverse groups

of secondary metabolites, and have multiple functions, such as carotenoid biosyn-

thesis, and protection from abiotic stresses (Heldt and Heldt, 2005). Although

possible explanations of our predictions are given, it is still unclear why the MVA

pathway predominates over the MEP pathway for the IPP biosynthesis under

drought conditions. Incorporating additional information of secondary metabolism

into the model constraints could greatly improve the accuracy of the model predic-

tions.

Lastly, numerous evidences suggested that light quality alters the accumulation

of flavonoids which are important secondary metabolites and important determi-

nants of fruit quality (Fu et al., 2016; Miao et al., 2016). Further curating the MEP

and MVA pathways in the metabolic model, and incorporating condition-specific

data will provide more insights into metabolic pathways influenced by light quality.

However, despite these limitations discussed above, this study provides a plat-

form for investigating metabolic activities under various growth conditions.

Chapter 6

106

6.3 Future perspective

Computational modelling is an iterative process with the aim of further improving

the model for better reflecting the biological phenomena in nature. Consequently,

in the future, it would be very interesting to experimentally test the hypotheses gen-

erated in this thesis, in particular, the results presented in Chapter 4. In addition to

experimental validations, the tasks that would be interesting and worthwhile pursu-

ing are as follows:

Improving gene-protein-reaction associations

Advances in high-throughput technologies have led to the generation of massive

amounts of omics data that must be interpreted in a biological meaningful way. With

the aid of metabolic models, the gene-protein-reactions (GPR) associations provide

the means to integrate and contextualize omics data such as proteomic and tran-

scriptomic data. In a recent study of rice, omics data integration and modelling was

used to explore the light-specific transcriptional signatures of cellular metabolism

(Lakshmanan et al., 2015). Transcriptomic and proteomic data has also been inte-

grated to a maize leaf genome-scale model as constraints for predicting metabolic

fluxes under different nitrogen conditions (Simons et al., 2014). In the last few

years, several computational methods have been developed to integrate omics data

into metabolic models to produce improved predictions of flux distribution (Kim

and Lun, 2014; Machado and Herrgård, 2014; Saha et al., 2014). These methods

depend extensively on the accuracy with which genes have been mapped to the bi-

ochemical reactions in the metabolic networks.

The determination of the GPR associated with each metabolic reaction is typi-

cally conducted by a laborious and time-consuming literature research. In the

presented tomato model, the GPR associations are included by listing gene(s) and

enzyme(s) involved in a specific reaction, but not refined with reliable evidence for

gen-reaction associations. In the future, it is necessary to improve GPR associations

in a new model for utilizing available tomato omics profiles to either produce new

tissue-specific models or constrain the behaviors of individual reaction in a meta-

bolic network.

Developing a fruit-specific model for tomato

General discussion

107

In the present work, the main focus was to study how the metabolic activity of leaf

cells varies in response to different growth condtions. Given that tomato is the most

studied model system of fleshy fruit ripening, and it contains several classes of an-

tioxidants such as carotenoids and phenolic compounds, it would be highly

interesting to model the metabolic behaviour of tomato fruit in future studies. To do

so, the approach proposed in other studies can be utilized to develop a fruit-specific

model (Grafahrend-Belau et al., 2009; Jerby et al., 2010; Mintz-Oron et al., 2012;

Seaver et al., 2015). Moreover, a stoichiometric model at medium-scale has been

published for tomato fruit (Colombié et al., 2015), which provides a reference to

build a new fruit-specific model at a large/genome-scale.

Multi-tissue modelling at whole-plant level

Another extension of the current work is to develop multi-tissue models at whole-

plant level. In planta, most of its metabolic functions depend on interactions be-

tween different tissues and organs, for instance, source-sink interactions, in which

the balance between source and sink is important for plant growth and yield (Osorio

et al., 2014; Li et al., 2015). Therefore, a whole-plant level understanding of plant

metabolism is required. In this regard, efforts toward modelling whole plant metab-

olism have already been made in barley (Grafahrend-Belau et al., 2013) and

Arabidopsis (de Oliveira Dal’Molin et al., 2015), respectively (see a review paper

by Martins Conde et al. (2016)). Basically, the two presented whole-plant metabolic

models are developed by integrating multiple tissue-specific models, such as leaf

model and stem model. A whole-plant model framework for tomato can also be

developed by implementing such an approach. In this multi-tissue context, it will be

interesting to explore resource allocation (e.g., carbon) in leaf (source) and fruit

(sink) in response to different light conditions.

Modelling plant-pathogen metabolic interactions

Constraint-based metabolic modelling has also opened new horizons in studying

metabolic interactions between species, including microbe-microbe and host-mi-

crobe interactions. Yet, there exist a few of examples in the literature (Jain and

Srivastava, 2009; Raghunathan et al., 2009; Heinken et al., 2013; Greenblum et al.,

2013; Manor et al., 2014). For example, Stolyar et al. (2007) examined the

syntrophic interactions between Desulfovibrio vulgaris and Methanococcus

maripaludis by using constraint-based modelling. In the study of Heinken et al.

Chapter 6

108

(2013), the metabolic interactions between a gut microbe Bacteroides thetaiotaomi-

cron and a mouse host was investigated using the constraint-based modelling

framework. Taken together, all these studies allow us to systematically investigate

complex metabolic interactions between species.

In addition to abiotic stresses such as drought and high light, the plant growth,

its health and crop yield can also be severely affected by biotic stresses. Among

these pathogens, root-knot nematodes (RKNs; Meloidogyne spp.) are major agricul-

tural pests in tomato. It was reported that more than 27% yield losses in tomato is

caused by root-knot nematodes (Kaur et al., 2011). Therefore, unraveling the plant-

pathogen metabolic interactions is a promising aspect to understand the pathogenic

effect of pathogens on host metabolic capacity and crop improvement. Following

the approaches proposed in previous studies (Heinken et al., 2013; Jamshidi and

Raghunathan, 2015), the metabolic interactions between tomato and RKNs can also

be modelled. For this purpose, efforts should be made towards: (1) constructing a

genome-scale metabolic model (GSM) for RKNs based on its annotated genome

(Abad et al., 2008); (2) integrating the two GSMs of PKNs and tomato with the

addition of a joint external compartment allowing metabolite exchanges; (3) per-

forming FBA simulations by employing appropriate objective functions with the

imposed constraints.

109

References

Abad, P., Gouzy, J., Aury, J.-M., et al. (2008) Genome sequence of the metazoan

plant-parasitic nematode Meloidogyne incognita. Nat. Biotechnol., 26,

909–915.

Affek, H.P. and Yakir, D. (2002) Protection by isoprene against singlet oxygen in

leaves. Plant Physiol., 129, 269–277.

Allen, J. (2002) Photosynthesis of ATP-electrons, proton pumps, rotors, and poise.

Cell, 110, 273–276.

Allen, J.F. (2003) Cyclic, pseudocyclic and noncyclic photophosphorylation: new

links in the chain. Trends Plant Sci., 8, 15–19.

Alonso, A.P., Goffman, F.D., Ohlrogge, J.B. and Shachar-Hill, Y. (2007) Carbon

conversion efficiency and central metabolic fluxes in developing sunflower

(Helianthus annuus L.) embryos. Plant J., 52, 296–308.

Antongiovanni, M. and Sargentini, C. (1991).Variability in chemical composition

of straws. Opt. Méditerranéen. Série-Sémin. 16, 49–53.

Antoniewicz, M.R. (2015) Methods and advances in metabolic flux analysis: a

mini-review. J. Ind. Microbiol. Biotechnol., 42, 317–325.

Araújo, W.L., Nunes-Nesi, A., Osorio, S., et al. (2011) Antisense inhibition of the

iron-sulphur subunit of succinate dehydrogenase enhances photosynthesis

and growth in tomato via an organic acid-mediated effect on stomatal aper-

ture. Plant Cell, 23, 600–627.

Araújo, W.L., Nunes-Nesi, A., Nikoloski, Z., Sweetlove, L.J. and Fernie, A.R. (2012a) Metabolic control and regulation of the tricarboxylic acid cycle in

photosynthetic and heterotrophic plant tissues. Plant Cell Environ., 35, 1–

21.

Araújo, W.L., Tohge, T., Osorio, S., et al. (2012b) Antisense inhibition of the 2-

oxoglutarate dehydrogenase complex in tomato demonstrates its im-

portance for plant respiration and during leaf senescence and fruit

maturation. Plant Cell, 24, 2328–2351.

References

110

Arnold, A. and Nikoloski, Z. (2014) Bottom-up metabolic reconstruction of Ara-

bidopsis and its application to determining the metabolic costs of enzyme

production. Plant Physiol., 165, 1380–1391.

Arnold, A., Sajitz-Hermstein, M. and Nikoloski, Z. (2015) Effects of varying ni-

trogen sources on amino acid synthesis costs in Arabidopsis thaliana under

different light and carbon-source conditions. PLoS One, 10, e0116536.

Atanasov, A.G., Waltenberger, B., Pferschy-Wenzig, E.-M., et al. (2015) Dis-

covery and resupply of pharmacologically active plant-derived natural

products: A review. Biotechnol. Adv., 33, 1582–1614.

Back, K., He, S., Kim, K.U. and Shin, D.H. (1998) Cloning and bacterial expres-

sion of sesquiterpene cyclase, a key branch point enzyme for the synthesis

of sesquiterpenoid phytoalexin capsidiol in UV-challenged leaves of Cap-

sicum annuum. Plant Cell Physiol., 39, 899–904.

Baghalian, K., Hajirezaei, M.-R. and Schreiber, F. (2014) Plant metabolic mod-

eling: achieving new insight into metabolism and metabolic engineering.

Plant Cell, 26, 3847–3866.

Bailleul, B., Berne, N., Murik, O., et al. (2015) Energetic coupling between plas-

tids and mitochondria drives CO2 assimilation in diatoms. Nature, 524,

366–369.

Basler, G., Küken, A., Fernie, A.R. and Nikoloski, Z. (2016) Photorespiratory

bypasses lead to increased growth in Arabidopsis thaliana: are predictions

consistent with experimental evidence? Synth. Biol., 31.

Bassel, G.W., Gaudinier, A., Brady, S.M., Hennig, L., Rhee, S.Y. and De Smet,

I. (2012) Systems analysis of plant functional, transcriptional, physical in-

teraction, and metabolic networks. Plant Cell, 24, 3859–3875.

Bauwe, H., Hagemann, M., Kern, R. and Timm, S. (2012) Photorespiration has

a dual origin and manifold links to central metabolism. Curr. Opin. Plant

Biol., 15, 269–275.

Becker, S.A., Feist, A.M., Mo, M.L., Hannum, G., Palsson, B.Ø. and Herrgard,

M.J. (2007) Quantitative prediction of cellular metabolism with constraint-

based models: the COBRA Toolbox. Nat. Protoc., 2, 727–738.

Blanco, N.E., Guinea-Díaz, M., Whelan, J. and Strand, Å. (2014) Interaction

between plastid and mitochondrial retrograde signalling pathways during

changes to plastid redox status. Philos. Trans. R. Soc. B., 369, 20130231.

References

111

Bonarius, H.P.J., Schmid, G. and Tramper, J. (1997) Flux analysis of underde-

termined metabolic networks: the quest for the miss in constraints. Trends

Biotechnol., 15, 308–314.

Bonnett, G.D. and Incoll, L.D. (1993a) Effects on the stem of winter barley of

manipulating the source and sink during grain-filling I. Changes in accu-

mulation and loss of mass from internodes.. J. Exp. Bot., 44, 75–82.

Bonnett, G.D. and Incoll, L.D. (1993b) Effects on the stem of winter barley of

manipulating the source and sink during grain-filling II. Changes in the

composition of water-souble carbohydrate of internodes. J. Exp. Bot., 44,

83–91.

Bordbar, A., Monk, J.M., King, Z.A. and Palsson, B.O. (2014) Constraint-based

models predict metabolic and associated cellular functions. Genetics, 15,

107–120.

Boyer, J.S. (1982) Plant productivity and environment. Science, 218, 443–448.

Bray, E.A. (1988) Drought- and ABA-induced changes in polypeptide and mRNA

accumulation in tomato leaves. Plant Physiol., 88, 1210–1214.

Brazaitytè, A., Duchovskis, P., Urbonavičiūtė, A., et al. (2009) The effects of

light-emitting diodes lighting on cucumber transplants and after effect on

yield. Zemdirbyste-Agriculture, 96,102–118.

Brocker, C., Vasiliou, M., Carpenter, S., et al. (2013) Aldehyde dehydrogenase

(ALDH) superfamily in plants: gene nomenclature and comparative ge-

nomics. Planta, 237, 189–210.

Bunik, V.I. and Fernie, A.R. (2009) Metabolic control exerted by the 2-oxoglu-

tarate dehydrogenase reaction: a cross-kingdom comparison of the

crossroad between energy production and nitrogen assimilation. Biochem.

J., 422, 405–421.

Carrari, F., Nunes-Nesi, A., Gibon, Y., Lytovchenko, A., Loureiro, M.E. and

Fernie, A.R. (2003) Reduced expression of aconitase results in an enhanced

rate of photosynthesis and marked shifts in carbon partitioning in illumi-

nated leaves of wild species tomato. Plant Physiol., 133, 1322–1335.

Caspi, R., Foerster, H., Fulcher, C.A., et al. (2008) The MetaCyc Database of

metabolic pathways and enzymes and the BioCyc collection of Path-

way/Genome Databases. Nucleic Acids Res., 36, D623–D631.

References

112

Chang, R.L., Ghamsari, L., Manichaikul, A., et al. (2011) Metabolic network

reconstruction of Chlamydomonas offers insight into light-driven algal me-

tabolism. Mol. Syst. Biol., 7, 518.

Chaves, M.M., Flexas, J. and Pinheiro, C. (2009) Photosynthesis under drought

and salt stress: regulation mechanisms from whole plant to cell. Ann. Bot.,

103, 551–560.

Cheung, C.Y.M., Poolman, M.G., Fell, D.A., Ratcliffe, R.G. and Sweetlove, L.J. (2014) A diel flux balance model captures interactions between light and

dark metabolism during day-night cycles in C3 and Crassulacean acid me-

tabolism leaves. Plant Physiol., 165, 917–929.

Cheung, C.Y.M., Ratcliffe, R.G. and Sweetlove, L.J. (2015) A method of ac-

counting for enzyme costs in flux balance analysis reveals alternative

pathways and metabolite stores in an illuminated Arabidopsis leaf. Plant

Physiol., 169, 1671–1682.

Cheung, C.Y.M., Williams, T.C.R., Poolman, M.G., Fell, D.A., Ratcliffe, R.G.

and Sweetlove, L.J. (2013) A method for accounting for maintenance costs

in flux balance analysis improves the prediction of plant cell metabolic phe-

notypes under stress conditions. Plant J., 75, 1050–1061.

Chung, B.K.-S., Lakshmanan, M., Klement, M., Mohanty, B. and Lee, D.-Y. (2013) Genome-scale in silico modeling and analysis for designing syn-

thetic terpenoid-producing microbial cell factories. Chem. Eng. Sci., 103,

100–108.

Chung, B.K.S. and Lee, D.-Y. (2009) Flux-sum analysis: a metabolite-centric ap-

proach for understanding the metabolic network. BMC Syst. Biol., 3, 117.

Collakova, E., Yen, J.Y. and Senger, R.S. (2012) Are we ready for genome-scale

modeling in plants? Plant Sci., 191–192, 53–70.

Colombié, S., Nazaret, C., Bénard, C., et al. (2015) Modelling central metabolic

fluxes by constraint-based optimization reveals metabolic reprogramming

of developing Solanum lycopersicum (tomato) fruit. Plant J., 81, 24–39.

Considine, M.J., Holtzapffel, R.C., Day, D.A., Whelan, J. and Millar, A.H. (2002) Molecular distinction between alternative oxidase from monocots

and dicots. Plant Physiol., 129, 949–953.

Cornic, G. (2000) Drought stress inhibits photosynthesis by decreasing stomatal

aperture – not by affecting ATP synthesis. Trends Plant Sci., 5, 187–188.

References

113

Cornish, K. and Zeevaart, J.A. (1985) Abscisic acid accumulation by roots of

Xanthium strumarium L. and Lycopersicon esculentum Mill. in relation to

water stress. Plant Physiol., 79, 653–658.

Coschigano, K.T., Melo-Oliveira, R., Lim, J. and Coruzzi, G.M. (1998) Ara-

bidopsis gls mutants and distinct Fd-GOGAT genes. Implications for

photorespiration and primary nitrogen assimilation. Plant Cell, 10, 741–

752.

Dani, K.G.S., Jamie, I.M., Prentice, I.C. and Atwell, B.J. (2014) Increased ratio

of electron transport to net assimilation rate supports elevated isoprenoid

emission rate in eucalypts under drought. Plant Physiol., 166, 1059–1072.

DeBolt, S., Scheible, W.-R., Schrick, K., et al. (2009) Mutations in UDP-

Glucose:sterol glucosyltransferase in Arabidopsis cause transparent testa

phenotype and suberization defect in seeds. Plant Physiol., 151, 78–87.

de Oliveira Dal’Molin, C.G., Quek, L.-E., Palfreyman, R.W., Brumbley, S.M.

and Nielsen, L.K. (2010a) AraGEM, a genome-scale reconstruction of the

primary metabolic network in Arabidopsis. Plant Physiol., 152, 579–589.

de Oliveira Dal’Molin, C.G., Quek, L.-E., Palfreyman, R.W., Brumbley, S.M.

and Nielsen, L.K. (2010b) C4GEM, a genome-scale metabolic model to

study C4 plant metabolism. Plant Physiol., 154, 1871–1885.

de Oliveira Dal’Molin, C.G. and Nielsen, L.K. (2013) Plant genome-scale meta-

bolic reconstruction and modelling. Curr. Opin. Biotechnol., 24, 271–277.

de Oliveira Dal’Molin, C.G, Quek, L.-E., Saa, P.A. and Nielsen, L.K. (2015) A

multi-tissue genome-scale metabolic modeling framework for the analysis

of whole plant systems. Front. Plant Sci., 6, 4.

Dharmawardhana, P., Ren, L., Amarasinghe, V., Monaco, M., Thomason, J.,

Ravenscroft, D., McCouch, S., Ware, D. and Jaiswal, P. (2013) A ge-

nome scale metabolic network for rice and accompanying analysis of

tryptophan, auxin and serotonin biosynthesis regulation under biotic stress.

Rice, 6, 15.

Dikicioglu, D., Kırdar, B. and Oliver, S.G. (2015) Biomass composition: the “el-

ephant in the room” of metabolic modelling. Metabolomics, 11, 1690–1701.

Döermann, P., Hoffmann-Benning, S., Balbo, I. and Be ning, C. (1995) Isolation

and characterization of an Arabidopsis mutant deficient in the thylakoid li-

pid digalactosyl diacylglycerol. Plant Cell, 7, 1801–1810.

References

114

Douce, R. and Neuburger, M. (1999) Biochemical dissection of photorespiration.

Curr. Opin. Plant Biol., 2, 214–222.

Dixon, R.A. (2005) A two-for-one in tomato nutritional enhancement. Nat. Biotech-

nol., 23, 825–826.

Durot, M., Bourguignon, P.-Y. and Schachter, V. (2009) Genome-scale models

of bacterial metabolism: reconstruction and applications. FEMS Microbiol.

Rev., 33, 164–190.

Eaton-Rye, J.J., Tripathy, B.C. and Sharkey, T.D. (2011) Photosynthesis: plas-

tid biology, energy conversion and carbon assimilation, Springer Science

& Business Media.

Ebrahim, A., Lerman, J.A., Palsson, B.O. and Hyduke, D.R. (2013) COBRApy:

COnstraints-based reconstruction and analysis for Python. BMC Syst. Biol.,

7, 74.

Edwards, J.M., Roberts, T.H. and Atwell, B.J. (2012) Quantifying ATP turnover

in anoxic coleoptiles of rice (Oryza sativa) demonstrates preferential allo-

cation of energy to protein synthesis. J. Exp. Bot., ers114.

El-Sayed, H.E.A. (2013) Exogenous application of ascorbic acid for improve ger-

mination, growth, water relations, organic and inorganic components in

tomato (Lycopersicon esculentum Mill.) plant under salt-stress. NY Sci. J.,

6, 123–139.

Engel, N., Daele, K. van den, Kolukisaoglu, Ü., Morgenthal, K., Weckwerth,

W., Pärnik, T., Keerberg, O. and Bauwe, H. (2007) Deletion of glycine

decarboxylase in Arabidopsis is lethal under nonphotorespiratory condi-

tions. Plant Physiol., 144, 1328–1335.

Fan, X.-X., Xu, Z.-G., Liu, X.-Y., Tang, C.-M., Wang, L.-W. and Han, X. (2013)

Effects of light intensity on the growth and leaf development of young to-

mato plants grown under a combination of red and blue light. Sci. Hortic.,

153, 50–55.

Feist, A.M., Henry, C.S., Reed, J.L., Krummenacker, M., Joyce, A.R., Karp,

P.D., Broadbelt, L.J., Hatzimanikatis, V. and Palsson, B.Ø. (2007) A

genome-scale metabolic reconstruction for Escherichia coli K-12 MG1655

that accounts for 1260 ORFs and thermodynamic information. Mol. Syst.

Biol., 3, 121.

References

115

Feist, A.M., Herrgård, M.J., Thiele, I., Reed, J.L. and Palsson, B.Ø. (2009) Re-

construction of biochemical networks in microorganisms. Nat. Rev.

Microbiol., 7, 129–143.

Feist, A.M. and Palsson, B.O. (2010) The biomass objective function. Curr. Opin.

Microbiol., 13, 344–349.

Field, K.J. and Lake, J.A. (2011) Environmental metabolomics links genotype to

phenotype and predicts genotype abundance in wild plant populations.

Physiol. Plant., 142, 352–360.

Flexas, J., Diaz-Espejo, A., Galmés, J., Kaldenhoff, R., Medrano, H. and Ribas-

Carbo, M. (2007) Rapid variations of mesophyll conductance in response

to changes in CO2 concentration around leaves. Plant Cell Environ., 30,

1284–1298.

Food and Agriculture Organization of the United Nations (FAO) (2014) Avail-

able online: http://faostat3.fao.org/home/E.

Flügge, U.-I. and Gao, W. (2005) Transport of isoprenoid intermediates across

chloroplast envelope membranes. Plant Biol. (Stuttgart), 7, 91–97.

Fortunati, A., Barta, C., Brilli, F., Centritto, M., Zimmer, I., Schnitzler, J.-P.

and Loreto, F. (2008) Isoprene emission is not temperature-dependent dur-

ing and after severe drought-stress: a physiological and biochemical

analysis. Plant J., 55, 687–697.

Foyer, C.H., Bloom, A.J., Queval, G. and Noctor, G. (2009) Photorespiratory

metabolism: genes, mutants, energetics, and redox signaling. Annu. Rev.

Plant Biol., 60, 455–484.

Fu, B., Ji, X., Zhao, M., He, F., Wang, X., Wang, Y., Liu, P. and Niu, L. (2016)

The influence of light quality on the accumulation of flavonoids in tobacco

(Nicotiana tabacum L.) leaves. J. Photochem. Photobiol. B, 162, 544–549.

Gevorgyan, A., Poolman, M.G. and Fell, D.A. (2008) Detection of stoichiometric

inconsistencies in biomolecular models. Bioinformatics, 24, 2245–2251.

Grafahrend-Belau, E., Junker, A., Eschenröder, A., Müller, J., Schreiber, F.

and Junker, B.H. (2013) Multiscale metabolic modeling: dynamic flux

balance analysis on a whole-plant scale. Plant Physiol., 163, 637–647.

References

116

Grafahrend-Belau, E., Schreiber, F., Koschützki, D. and Junker, B.H. (2009)

Flux balance analysis of barley seeds: a computational approach to study

systemic properties of central metabolism. Plant Physiol., 149, 585–598.

Greenblum, S., Chiu, H.-C., Levy, R., Carr, R. and Borenstein, E. (2013) To-

wards a predictive systems-level model of the human microbiome:

progress, challenges, and opportunities. Curr. Opin. Biotechnol., 24, 810–

820.

Griffin, K.L. and Turnbull, M.H. (2012) Out of the light and into the dark: post-

illumination respiratory metabolism. New Phytol., 195, 4–7.

Guinn, G. (1966) Extraction of nucleic acids from lyophilized plant material. Plant

Physiol., 41, 689–695.

Hamilton, J.J. and Reed, J.L. (2014) Software platforms to facilitate reconstruct-

ing genome-scale metabolic networks. Environ. Microbiol., 16, 49–59.

Hanson, A.D. and Roje, S. (2001) One-carbon metabolism in higher plants. Annu.

Rev. Plant Biol., 52, 119–137.

Harris, M.J. and Outlaw, W.H. (1991) Rapid adjustment of guard-cell abscisic

acid levels to current leaf-water status 1. Plant Physiol., 95, 171–173.

Hartman, H.B., Fell, D.A., Rossell, S., et al. (2014a) Identification of potential

drug targets in Salmonella enterica sv. Typhimurium using metabolic mod-

elling and experimental validation. Microbiology, 160, 1252–1266.

Haupt-Herting, S., Klug, K. and Fock, H.P. (2001) A new approach to measure

gross CO2 fluxes in leaves. Gross CO2 assimilation, photorespiration, and

mitochondrial respiration in the light in tomato under drought stress. Plant

Physiol., 126, 388–396.

Hay, J. and Schwender, J. (2011) Metabolic network reconstruction and flux var-

iability analysis of storage synthesis in developing oilseed rape (Brassica

napus L.) embryos. Plant J., 67, 526–541.

Heinken, A., Sahoo, S., Fleming, R.M.T. and Thiele, I. (2013) Systems-level

characterization of a host-microbe metabolic symbiosis in the mammalian

gut. Gut Microbes, 4, 28–40.

Heldt, H.-W. and Heldt, F. (2005) Plant Biochemistry, Academic Press.

References

117

Hernández, J.M.G.M., Castignolles, P., Gidley, M.J., Myers, A.M. and Gilbert,

R.G. (2008) Mechanistic investigation of a starch-branc hing enzyme using

hydrodynamic volume SEC analysis. Biomacromolecules, 9, 954–965.

Hoefnagel, M.H.N., Atkin, O.K. and Wiskich, J.T. (1998) Interdependence be-

tween chloroplasts and mitochondria in the light and the dark. Biochim.

Biophys. Acta, Bioenerg., 1366, 235–255.

Holzhütter, H.-G. (2004) The principle of flux minimization and its application to

estimate stationary fluxes in metabolic networks: flux minimization. Eur. J.

Biochem., 271, 2905–2922.

Hounsome, N., Hounsome, B., Tomos, D. and Edwards-Jones, G. (2008) Plant

metabolites and nutritional quality of vegetables. J. Food Sci., 73, R48-65.

Hucka, M., Finney, A., Sauro, H.M., et al. (2003) The systems biology markup

language (SBML): a medium for representation and exchange of biochem-

ical network models. Bioinformatics, 19, 524–531.

Imam, S., Fitzgerald, C.M., Cook, E.M., Donohue, T.J. and Noguera, D.R. (2015) Quantifying the effects of light intensity on bioproduction and

maintenance energy during photosynthetic growth of Rhodobacter

sphaeroides. Photosynth. Res., 123, 167–182.

Ip, K., Donoghue, N., Kim, M.K. and Lun, D.S. (2014) Constraint-based model-

ing of heterologous pathways: application and experimental demonstration

for overproduction of fatty acids in Escherichia coli. Biotechnol. Bioeng.,

111, 2056–2066.

Jamshidi, N. and Raghunathan, A. (2015) Cell scale host-pathogen modeling: an-

other branch in the evolution of constraint-based methods. Infect. Dis.,

1032.

Jardine, K., Chambers, J., Alves, E.G., et al. (2014) Dynamic balancing of iso-

prene carbon sources reflects photosynthetic and photorespiratory

responses to temperature stress. Plant Physiol., 166, 2051–2064.

Jerby, L., Shlomi, T. and Ruppin, E. (2010) Computational reconstruction of tis-

sue-specific metabolic models: application to human liver metabolism. Mol.

Syst. Biol., 6, 401.

Jordan, D.B. and Ogren, W.L. (1984) The CO2/O2 specificity of ribulose 1,5-

bisphosphate carboxylase/oxygenase. Planta, 161, 308–313.

References

118

Josse, E.-M., Simkin, A.J., Gaffé, J., Labouré, A.-M., Kuntz, M. and Carol, P. (2000) A plastid terminal oxidase associated with carotenoid desaturation

during chromoplast differentiation. Plant Physiol., 123, 1427–1436.

Juliano, B.O. (1985) “Rice hull and rice straw,” in Rice: Chemistry and Technol-

ogy, ed B. O. Julano (St. Paul, MN: The American Asso ciation of Cereal

Chemists Incorporated), 699.

Junker, B.H., Lonien, J., Heady, L.E., Rogers, A. and Schwender, J. (2007) Par-

allel determination of enzyme activities and in vivo fluxes in Brassica napus

embryos grown on organic or inorganic nitrogen source. Phytochemistry,

68, 2232–2242.

Kahlau, S. and Bock, R. (2008) Plastid transcriptomics and translatomics of tomato

fruit development and chloroplast-to-chromoplast differentiation: chromo-

plast gene expression largely serves the production of a single protein. Plant

Cell, 20, 856–874.

Kanehisa, M. and Goto, S. (2000) KEGG: Kyoto Encyclopedia of Genes and Ge-

nomes. Nucleic Acids Res., 28, 27–30.

Kargul, J., Nield, J. and Barber, J. (2003) Three-dimensional reconstruction of a

light-harvesting complex I-photosystem I (LHCI-PSI) supercomplex from

the green alga Chlamydomonas reinhardtii. Insights into light harvesting for

PSI. J. Biol. Chem., 278, 16135–16141.

Karp, P.D., Ouzounis, C.A., Moore-Kochlacs, C., et al. (2005) Expansion of the

BioCyc collection of pathway/genome databases to 160 genomes. Nucleic

Acids Res., 33, 6083–6089.

Katterman, F.R. and Ergle, D.R. (1966) RNA composition in cotton. Plant Phys-

iol., 41, 553–556.

Kauffman, K.J., Prakash, P. and Edwards, J.S. (2003) Advances in flux balance

analysis. Curr. Opin. Biotechnol., 14, 491–496.

Kaur, D.N., Sharma, S.K. and Sultan, M.S. (2011) Effect of different chemicals

on root knot nematode in seed beds of tomato. Plant Dis. Res., 26, 170-170.

Kim, H.-H., Wheeler, R.M., Sager, J.C., Yorio, N.C. and Goins, G.D. (2005)

Light-emitting diodes as an illumination source for plants: a review of re-

search at Kennedy Space Center. Habitation (Elmsford, NY, U. S.), 10, 71–

78.

References

119

Kim, J., Fabris, M., Baart, G., Kim, M.K., Goossens, A., Vyverman, W.,

Falkowski, P.G. and Lun, D.S. (2016) Flux balance analysis of primary

metabolism in the diatom Phaeodactylum tricornutum. Plant J., 85, 161–

176.

Kim, M.K. and Lun, D.S. (2014) Methods for integration of transcriptomic data in

genome-scale metabolic models. Comput. Struct. Biotechnol. J., 11, 59–65.

Kim, T., Dreher, K., Nilo-Poyanco, R., et al. (2015) Patterns of metabolite

changes identified from large-scale gene perturbations in Arabidopsis using

a genome-scale metabolic network. Plant Physiol., 167, 1685–1698.

Kim, T.Y., Sohn, S.B., Kim, Y.B., Kim, W.J. and Lee, S.Y. (2012) Recent ad-

vances in reconstruction and applications of genome-scale metabolic

models. Curr. Opin. Biotechnol., 23, 617–623.

Klamt, S., Schuster, S. and Gilles, E.D. (2002) Calculability analysis in underde-

termined metabolic networks illustrated by a model of the central

metabolism in purple nonsulfur bacteria. Biotechnol. Bioeng., 77, 734–751.

Klee, H.J. and Giovannoni, J.J. (2011) Genetics and control of tomato fruit ripen-

ing and quality attributes. Annu. Rev. Genet., 45, 41–59.

Koussa, J., Chaiboonchoe, A. and Salehi-Ashtiani, K. (2014) Computational ap-

proaches for microalgal biofuel optimization: a review. BioMed Res. Int.,

2014, 649453.

Kramer, D.M. and Evans, J.R. (2011) The importance of energy balance in im-

proving photosynthetic productivity. Plant Physiol., 155, 70–78.

Kruger, N.J. and Ratcliffe, R.G. (2009) Insights into plant metabolic networks

from steady-state metabolic flux analysis. Biochimie, 91, 697–702.

Kuc, J. (1995) Phytoalexins, stress metabolism, and disease resistance in plants.

Annu. Rev. Phytopathol., 33, 275–297.

Kwon, Y.W. and Soh, C.H. (1985) Characteristics of nuclear DNA in rice roots of

Japonica and Indica × Japonicavarieties. Agric. Res. Seoul Natl. Univ. 10,

63–68.

Lakshmanan, M., Koh, G., Chung, B.K.S. and Lee, D.-Y. (2014) Software ap-

plications for flux balance analysis. Brief. Bioinform., 15, 108–122.

References

120

Lakshmanan, M., Lim, S.-H., Mohanty, B., Kim, J.K., Ha, S.-H. and Lee, D.-

Y. (2015) Unraveling the light-specific metabolic and regulatory signatures

of rice through combined in silico modeling and multiomics analysis. Plant

Physiol., 169, 3002–3020.

Lakshmanan, M., Mohanty, B. and Lee, D.-Y. (2013b) Identifying essential

genes/reactions of the rice photorespiration by in silico model-based analy-

sis. Rice, 6, 20.

Lakshmanan, M., Zhang, Z., Mohanty, B., Kwon, J.-Y., Choi, H.-Y., Nam, H.-

J., Kim, D.-I. and Lee, D.-Y. (2013a) Elucidating rice cell metabolism un-

der flooding and drought stresses using flux-based modeling and analysis.

Plant Physiol., 162, 2140–2150.

Leegood, R.C. (2007) A welcome diversion from photorespiration. Nat. Biotech-

nol., 25, 539–540.

Levering, J., Broddrick, J., Dupont, C.L., et al. (2016) Genome-scale model re-

veals metabolic basis of biomass partitioning in a model diatom. PloS One,

11, e0155038.

Li, T., Heuvelink, E. and Marcelis, L.F.M. (2015) Quantifying the source–sink

balance and carbohydrate content in three tomato cultivars. Front. Plant

Sci., 6.

Lichtenthaler, H.K., Schwender, J., Disch, A. and Rohmer, M. (1997) Biosyn-

thesis of isoprenoids in higher plant chloroplasts proceeds via a mevalonate-

independent pathway. Eur. J. Biochem., 400, 271–274.

Lieberman, M., Segev, O., Gilboa, N., Lalazar, A. and Levin, I. (2004) The to-

mato homolog of the gene encoding UV-damaged DNA binding protein 1

(DDB1) underlined as the gene that causes the high pigment-1 mutant phe-

notype. Theor. Appl. Genet., 108, 1574–1581.

Liepman, A.H. and Olsen, L.J. (2001) Peroxisomal alanine: glyoxylate ami-

notransferase (AGT1) is a photorespiratory enzyme with multiple substrates

in Arabidopsis thaliana. Plant J., 25, 487–498.

Linka, M. and Weber, A.P.M. (2005) Shuffling ammonia between mitochondria

and plastids during photorespiration. Trends Plant Sci., 10, 461–465.

Liu, Y., Roof, S., Ye, Z., Barry, C., Tuinen, A. van, Vrebalov, J., Bowler, C. and

Giovannoni, J. (2004) Manipulation of light signal transduction as a means

References

121

of modifying fruit nutritional quality in tomato. Proc. Natl. Acad. Sci. U. S.

A., 101, 9897–9902.

Llorente, B., D’Andrea, L., Ruiz-Sola, M.A., Botterweg, E., Pulido, P., Andilla,

J., Loza-Alvarez, P. and Rodriguez-Concepcion, M. (2016) Tomato fruit

carotenoid biosynthesis is adjusted to actual ripening progression by a light-

dependent mechanism. Plant J., 85, 107–119.

Loreto, F. and Schnitzler, J.-P. (2010) Abiotic stresses and induced BVOCs.

Trends Plant Sci., 15, 154–166.

Loreto, F. and Velikova, V. (2001) Isoprene produced by leaves protects the pho-

tosynthetic apparatus against ozone damage, quenches ozone products, and

reduces lipid peroxidation of cellular membranes. Plant Physiol., 127,

1781–1787.

Lytovchenko, A., Eickmeier, I., Pons, C., et al. (2011) Tomato fruit photosynthe-

sis is seemingly unimportant in primary metabolism and ripening but plays

a considerable role in seed development. Plant Physiol., 157, 1650–1663.

Maarleveld, T.R., Boele, J., Bruggeman, F.J. and Teusink, B. (2014) A data in-

tegration and visualization resource for the metabolic network of

Synechocystis sp. PCC 6803. Plant Physiol., 164, 1111-1121.

Machado, D. and Herrgård, M. (2014) Systematic evaluation of methods for in-

tegration of transcriptomic data into constraint-based models of

metabolism. PLoS Comput. Biol., 10, e1003580.

Mahadevan, R. and Schilling, C.H. (2003) The effects of alternate optimal solu-

tions in constraint-based genome-scale metabolic models. Metab. Eng., 5,

264–276.

Manor, O., Levy, R. and Borenstein, E. (2014) Mapping the Inner Workings of

the Microbiome: Genomic- and Metagenomic-Based Study of Metabolism

and Metabolic Interactions in the Human Microbiome. Cell Metab., 20,

742–752.

Martins Conde, P. do R., Sauter, T. and Pfau, T. (2016) Constraint based mod-

eling going multicellular. Front. Mol. Biosci., 3.

Masakapalli, S.K., Le Lay, P., Huddleston, J.E., Pollock, N.L., Kruger, N.J.

and Ratcliffe, R.G. (2010) Subcellular flux analysis of central metabolism

in a heterotrophic Arabidopsis cell Suspension using steady-state stable iso-

tope labeling. Plant Physiol., 152, 602–619.

References

122

McCloskey, D., Palsson, B.Ø. and Feist, A.M. (2013) Basic and applied uses of

genome-scale metabolic network reconstructions of Escherichia coli. Mol.

Syst. Biol., 9, 661.

Mechin, V., Thevenot, C., Le Guilloux, M., Prioul, J.L. and Dame val, C. (2007)

Developmental analysis of maize endosperm proteome suggests a pivotal

role for pyruvate orthophosphate dikinase. Plant Physiol. 143, 1203–1219.

Miao, L., Zhang, Y., Yang, X., Xiao, J., Zhang, H., Zhang, Z., Wang, Y. and

Jiang, G. (2016) Colored light-quality selective plastic films affect antho-

cyanin content, enzyme activities, and the expression of flavonoid genes in

strawberry (Fragaria × ananassa) fruit. Food Chem., 207, 93–100.

Mintz-Oron, S., Meir, S., Malitsky, S., Ruppin, E., Aharoni, A. and Shlomi, T. (2012) Reconstruction of Arabidopsis metabolic network models account-

ing for subcellular compartmentalization and tissue-specificity. Proc. Natl.

Acad. Sci. U. S. A., 109, 339–344.

Mohanty, B., Kitazumi, A., Cheung, C.Y.M., Lakshmanan, M., Reyes, B.G. de

los, Jang, I.-C. and Lee, D.-Y. (2016) Identification of candidate network

hubs involved in metabolic adjustments of rice under drought stress by in-

tegrating transcriptome data and genome-scale metabolic network. Plant

Sci., 242, 224–239.

Moll, B. and Levine, R.P. (1970) Characterization of a photosynthetic mutant

strain of Chlamydomonas reinhardii deficient in phosphoribulokinase activ-

ity. Plant Physiol., 46, 576–580.

Mooney, B.P., Miernyk, J.A., Michael Greenlief, C. and Thelen, J.J. (2006) Us-

ing quantitative proteomics of Arabidopsis roots and leaves to predict

metabolic activity. Physiol. Plant., 128, 237–250.

Morfopoulos, C., Sperlich, D., Peñuelas, J., et al. (2014) A model of plant iso-

prene emission based on available reducing power captures responses to

atmospheric CO₂. New Phytol., 203, 125–139.

Mouillon, J.M., Aubert, S., Bourguignon, J., Gout, E., Douce, R. and Rébeillé,

F. (1999) Glycine and serine catabolism in non-photosynthetic higher plant

cells: their role in C1 metabolism. Plant J., 20, 197–205.

Muller, F., Dijkhuis, D.J. and Heida, Y.S. (1970) On the relationship between

chemical composition and digestibility in vivo of roughages. Agric. Res.

Rep., 736, 1–27.

References

123

Mueller, L.A., Zhang, P. and Rhee, S.Y. (2003) AraCyc: a biochemical pathway

database for Arabidopsis. Plant Physiol., 132, 453–460.

Muneer, S., Kim, E.J., Park, J.S. and Lee, J.H. (2014) Influence of green, red and

blue light emitting diodes on multiprotein complex proteins and photosyn-

thetic activity under different light intensities in lettuce leaves (Lactuca

sativa L.). Int. J. Mol. Sci., 15, 4657–4670.

Muthuraj, M., Palabhanvi, B., Misra, S., Kumar, V., Sivalingavasu, K. and

Das, D. (2013) Flux balance analysis of Chlorella sp. FC2 IITG under pho-

toautotrophic and heterotrophic growth conditions. Photosynth. Res., 118,

167–179.

Nakashima, K., Yamaguchi-Shinozaki, K. and Shinozaki, K. (2014) The tran-

scriptional regulatory network in the drought response and its crosstalk in

abiotic stress responses including drought, cold, and heat. Front. Plant Sci.,

5.

Nakazono, M., Tsuji, H., Li, Y., Saisho, D., Arimura, S., Tsutsumi, N. and Hi-

rai, A. (2000) Expression of a gene encoding mitochondrial aldehyde

dehydrogenase in rice increases under submerged conditions. Plant Phys-

iol., 124, 587–598.

Nield, J., Kruse, O., Ruprecht, J., Fonseca, P. da, Büchel, C. and Barber, J. (2000) Three-dimensional structure of Chlamydomonas reinhardtii and

Synechococcus elongatus photosystem II complexes allows for comparison

of their oxygen-evolving complex organization. J. Biol. Chem., 275,

27940–27946.

Niemann, G.J., Pureveen, J.B.M., Eijkel, G.B., Poorter, H. and Boon, J.J.

(1995) Differential chemical allocation and plant adaptation: a Py-MS study

of 24 species differing in relative growth rate. Plant Soil, 175, 275–289.

Nielsen, J. (2003) It is all about metabolic fluxes. J. Bacteriol., 185, 7031–7035.

Nogales, J., Gudmundsson, S., Knight, E.M., Palsson, B.O. and Thiele, I. (2012)

Detailing the optimality of photosynthesis in cyanobacteria through systems

biology analysis. Proc. Natl. Acad. Sci. U. S. A., 109, 2678–2683.

Nookaew, I., Jewett, M.C., Meechai, A., Thammarongtham, C., Laoten, K.,

Cheevadhanarak, S., et al. (2008).The genome-scale metabolic model

ilN800 of Saccharomyces cerevisiae and its validation: a scaffold to query

lipid metabolism. BMC Syst. Biol., 2, 71.

References

124

Norton, G. (1989) “Nature and biosynthesis of storage proteins,” in Oil Crops of

the World, ed G. Röbbelen, R.K. Downey, and A. Ashri (New York, NY:

McGraw-Hill), 165–191.

Novák, L. and Loubiere, P. (2000) The metabolic network of Lactococcus lactis:

distribution of (14)C-labeled substrates between catabolic and anabolic

pathways. J. Bacteriol., 182, 1136–1143.

Nunes-Nesi, A., Carrari, F., Gibon, Y., et al. (2007) Deficiency of mitochondrial

fumarase activity in tomato plants impairs photosynthesis via an effect on

stomatal function. Plant J., 50, 1093–1106.

Oberhardt, M.A., Palsson, B.Ø. and Papin, J.A. (2009) Applications of genome-

scale metabolic reconstructions. Mol. Syst. Biol., 5, 320.

O’Carrigan, A., Babla, M., Wang, F., Liu, X., Mak, M., Thomas, R., Bellotti,

B. and Chen, Z.-H. (2014) Analysis of gas exchange, stomatal behaviour

and micronutrients uncovers dynamic response and adaptation of tomato

plants to monochromatic light treatments. Plant Physiol. Biochem., 82,

105–115.

OECD (2004) “Consensus document on compositional considerations for new va-

riety of barley (Hordeum vulgare): key food and nutrients,” in Series on the

Safety of Novel Foods and Feeds, ed OECD (Paris: OECD Environmental

Health and Safety Publications), 1–42.

Ohlrogge, J. and Browse, J. (1995) Lipid biosynthesis. Plant Cell, 7, 957–970.

Orth, J.D., Thiele, I. and Palsson, B.Ø. (2010) What is flux balance analysis? Nat.

Biotechnol., 28, 245–248.

Osorio, S., Ruan, Y.-L. and Fernie, A.R. (2014) An update on source-to-sink car-

bon partitioning in tomato. Front. Plant Sci., 5, 516.

Pegoraro, E., Rey, A., Greenberg, J., Harley, P., Grace, J., Malhi, Y. and Guen-

ther, A. (2004) Effect of drought on isoprene emission rates from leaves of

Quercus virginiana Mill. Atmos. Environ., 38, 6149–6156.

Penning de Vries, F.W., Brunsting, A.H. and Laar, H.H. van (1974) Products,

requirements and efficiency of biosynthesis: a quantitative approach. J.

Theor. Biol., 45, 339–377.

Pilalis, E., Chatziioannou, A., Thomasset, B. and Kolisis, F. (2011) An in silico

compartmentalized metabolic model of Brassica napus enables the systemic

References

125

study of regulatory aspects of plant central metabolism. Biotechnol. Bio-

eng., 108, 1673–1682.

Pinchuk, G.E., Hill, E.A., Geydebrekht, O.V., Ingeniis, J.D., Zhang, X., Oster-

man, A., et al. (2010) Constraint-based model of Shewanella oneidensis

MR-1 metabolism: a tool for data analysis and hypothesis generation. PLoS

Comput. Biol., 6:e1000822.

Pinheiro, C. and Chaves, M.M. (2011) Photosynthesis and drought: can we make

metabolic connections from available data? J. Exp. Bot., 62, 869–882.

Poolman, M.G. (2006) ScrumPy: metabolic modelling with Python. Syst. Biol.

(Stevenage), 153, 375.

Poolman, M.G., Kundu, S., Shaw, R. and Fell, D.A. (2013) Responses to Light

Intensity in a Genome-Scale Model of Rice Metabolism. Plant Physiol.,

162, 1060–1072.

Poolman, M.G., Miguet, L., Sweetlove, L.J. and Fell, D.A. (2009) A genome-

scale metabolic model of Arabidopsis and some of its properties. Plant

Physiol., 151, 1570–1581.

Poorter, H. and Bergkotte, M. (1992) Chemical composition of 24 wild species

differing in relative growth rate. Plant Cell Environ., 15, 221–229.

Pramanik, J. and Keasling, J.D. (1998) Effect of Escherichia coli biomass com-

position on central metabolic fluxes predicted by a stoichiometric model.

Biotechnol. Bioeng., 60, 230–238.

Pramanik, J. and Keasling, J.D. (1997) Stoichiometric model of Escherichia coli

metabolism: incorporation of growth-rate dependent biomass composition

and mechanistic energy requirements. Biotechnol. Bioeng., 56, 398–421.

Price, N.D., Reed, J.L. and Palsson, B.Ø. (2004) Genome-scale models of micro-

bial cells: evaluating the consequences of constraints. Nat. Rev. Microbiol.,

2, 886–897.

Puchałka, J., Oberhardt, M.A., Godinho, M., Bielecka, A., Regenhardt, D.,

Timmis, K.N., Papin, J.A. and Martins dos Santos, V.A.P. (2008) Ge-

nome-scale reconstruction and analysis of the Pseudomonas putida KT2440

metabolic network facilitates applications in biotechnology. PLoS Comput.

Biol., 4, e1000210.

References

126

Pyl, E.-T., Piques, M., Ivakov, A., Schulze, W., Ishihara, H., Stitt, M. and Sul-

pice, R. (2012) Metabolism and growth in Arabidopsis depend on the

daytime temperature but are temperature-compensated against cool nights.

Plant Cell, 24, 2443–2469.

Quick, W.P., Schurr, U., Scheibe, R., Schulze, E.D., Rodermel, S.R., Bogorad,

L. and Stitt, M. (1991) Decreased ribulose-1,5-bisphosphate carboxylase-

oxygenase in transgenic tobacco transformed with “antisense” rbcS: I.

Impact on photosynthesis in ambient growth conditions. Planta, 183, 542–

554.

Rademacher, T., Hausler, R.E., Hirsch, H.J., Zhang, L., Lipka, V., Weier, D.,

Kreuzaler, F. and Peterhansel, C. (2002) An engineered phosphoenolpy-

ruvate carboxylase redirects carbon and nitrogen flow in transgenic potato

plants. Plant J., 32, 25–39.

Radrich, K., Tsuruoka, Y., Dobson, P., Gevorgyan, A., Swainston, N., Baart,

G. and Schwartz, J.-M. (2010) Integration of metabolic databases for the

reconstruction of genome-scale metabolic networks. BMC Syst. Biol., 4,

114.

Raghunathan, A., Reed, J., Shin, S., Palsson, B. and Daefler, S. (2009) Con-

straint-based analysis of metabolic capacity of Salmonella typhimurium

during host–pathogen interaction. BMC Syst. Biol., 3, 38.

Reinders, A., Panshyshyn, J.A. and Ward, J.M. (2005) Analysis of transport ac-

tivity of Arabidopsis sugar alcohol permease homolog AtPLT5. J. Biol.

Chem., 280, 1594–1602.

Rischer, H., Häkkinen, S.T., Ritala, A., Seppänen-Laakso, T., Miralpeix, B.,

Capell, T., Christou, P. and Oksman-Caldentey, K.-M. (2013) Plant

cells as pharmaceutical factories. Curr. Pharm. Des., 19, 5640–5660.

Rodríguez-Concepción, M., Forés, O., Martinez-García, J.F., González, V.,

Phillips, M.A., Ferrer, A. and Boronat, A. (2004) Distinct light-mediated

pathways regulate the biosynthesis and exchange of isoprenoid precursors

during Arabidopsis seedling development. Plant Cell, 16, 144–156.

Roessner-Tunali, U., Hegemann, B., Lytovchenko, A., Carrari, F., Bruedigam,

C., Granot, D. and Fernie, A.R. (2003) Metabolic profiling of transgenic

tomato plants overexpressing hexokinase reveals that the influence of hex-

ose phosphorylation diminishes during fruit development. Plant Physiol.,

133, 84–99.

References

127

Rontein, D., Dieuaide-Noubhani, M., Dufourc, E.J., Raymond, P. and Rolin, D. (2002) The metabolic architecture of plant cells. Stability of central metab-

olism and flexibility of anabolic pathways during the growth cycle of

tomato cells. J. Biol. Chem., 277, 43948–43960.

Rügen, M., Bockmayr, A. and Steuer, R. (2015) Elucidating temporal resource

allocation and diurnal dynamics in phototrophic metabolism using condi-

tional FBA. Sci. Rep., 5, 15247.

Sager, J.C. and McFarlane, J.C. (1997) in Plant growth chamber handbook, Ra-

diation, eds Langhans R.W., Tibbitts T.W. (Iowa State Univ. Press: North

Central Region Research Publication No. 340, Iowa Agriculture and Home

Economics Experiment Station Special Report no. 99, Ames, IA), pp 1–29.

Saha, R., Chowdhury, A. and Maranas, C.D. (2014) Recent advances in the re-

construction of metabolic models and integration of omics data. Curr. Opin.

Biotechnol., 29, 39–45.

Saha, R., Suthers, P.F. and Maranas, C.D. (2011) Zea mays i RS1563: A com-

prehensive genome-scale metabolic reconstruction of maize metabolism.

PLoS One, 6, e21784.

Sánchez-Rodríguez, E., Rubio-Wilhelmi, M., Cervilla, L.M., Blasco, B., Rios,

J.J., Rosales, M.A., Romero, L. and Ruiz, J.M. (2010) Genotypic differ-

ences in some physiological parameters symptomatic for oxidative stress

under moderate drought in tomato plants. Plant Sci., 178, 30–40.

Schauer, N., Zamir, D. and Fernie, A.R. (2005) Metabolic profiling of leaves and

fruit of wild species tomato: a survey of the Solanum lycopersicum com-

plex. J. Exp. Bot., 56, 297–307.

Schomburg, I., Chang, A., Ebeling, C., Gremse, M., Heldt, C., Huhn, G. and

Schomburg, D. (2004) BRENDA, the enzyme database: updates and major

new developments. Nucleic Acids Res., 32, D431–D433.

Schuetz, R., Kuepfer, L. and Sauer, U. (2007) Systematic evaluation of objective

functions for predicting intracellular fluxes in Escherichia coli. Mol. Syst.

Biol., 3.

Schwab, W. (2003) Metabolome diversity: too few genes, too many metabolites?

Phytochemistry, 62, 837–849.

Schwarte, S. and Bauwe, H. (2007) Identification of the photorespiratory 2-phosp

References

128

hoglycolate phosphatase, PGLP1, in Arabidopsis. Plant Physiol., 144,

1580–1586.

Schwender, J., Goffman, F., Ohlrogge, J.B. and Shachar-Hill, Y. (2004) Rubisco

without the Calvin cycle improves the carbon efficiency of developing

green seeds. Nature, 432, 779–782.

Schwender, J. and Hay, J.O. (2012) Predictive modeling of biomass component

tradeoffs in Brassica napus developing oilseeds based on in silico manipu-

lation of storage metabolism. Plant Physiol., 160, 1218–1236.

Schwender, J. and Ohlrogge, J.B. (2002) Probing in vivo metabolism by stable

isotope labeling of storage lipids and proteins in developing Brassica napus

embryos. Plant Physiol., 130, 347–361.

Schwender, J., Shachar-Hill, Y. and Ohlrogge, J.B. (2006) Mitochondrial metab-

olism in developing embryos of Brassica napus. J. Biol. Chem., 281, 34040–

34047.

Seaver, S.M.D., Bradbury, L.M.T., Frelin, O., Zarecki, R., Ruppin, E., Hanson,

A.D. and Henry, C.S. (2015) Improved evidence-based genome-scale met-

abolic models for maize leaf, embryo, and endosperm. Front. Plant Sci., 6,

142.

Sharkey, T.D., Wiberley, A.E. and Donohue, A.R. (2008) Isoprene emission from

plants: why and how. Ann. Bot., 101, 5–18.

Sharrock, R.A. and Clack, T. (2002) Patterns of expression and normalized levels

of the five Arabidopsis phytochromes. Plant Physiol., 130, 442–456.

Sheen, S.J. (1983) Biomass and chemical composition of tobacco plants under high

density growth. Beitr. Tabakforsch. Int., 12, 35–42.

Shioi, Y. and Sasa, T. (1984) Chlorophyll formation in the YG-6 mutant of chlo-

rella regularis: spectral characterization of protochlorophyllide

phototransformation. Plant Cell Physiol., 25, 139–149.

Simons, M., Saha, R., Amiour, N., et al. (2014) Assessing the metabolic impact

of nitrogen availability using a compartmentalized maize leaf genome-scale

model. Plant Physiol., 166, 1659–1674.

Skillman, J.B. (2008) Quantum yield variation across the three pathways of photo-

synthesis: not yet out of the dark. J. Exp. Bot., 59, 1647–1661.

References

129

Smallbone, K. and Simeonidis, E. (2009) Flux balance analysis: a geometric per-

spective. J. Theor. Biol., 258, 311–315.

Smirnoff, N. (1993) The role of active oxygen in the response of plants to water

deficit and desiccation. New Phytol., 125, 27–58.

Smith, A.M. and Stitt, M. (2007) Coordination of carbon supply and plant growth.

Plant Cell Environ., 30, 1126–1149.

Spector, W.S. (1956) Handbook of Biological Data. Washington, DC:National

Academy of Science.

Spielbauer, G., Margl, L., Hannah, L.C., Römisch, W., Ettenhuber, C., Bacher,

A., Gierl, A., Eisenreich, W. and Genschel, U. (2006) Robustness of cen-

tral carbohydrate metabolism in developing maize kernels. Phytochemistry,

67, 1460–1475.

Sterman, J.D. (2002) All models are wrong: reflections on becoming a systems

scientist. Syst. Dyn. Rev., 18, 501–531.

Stolyar, S., Van Dien, S., Hillesland, K.L., Pinel, N., Lie, T.J., Leigh, J.A. and

Stahl, D.A. (2007) Metabolic modeling of a mutualistic microbial commu-

nity. Mol. Syst. Biol., 3, 92.

Story, E.N., Kopec, R.E., Schwartz, S.J. and Harris, G.K. (2010) An update on

the health effects of tomato lycopene. Annu. Rev. Food Sci. Technol., 1,

189–210.

Sulpice, R., Nikoloski, Z., Tschoep, H., et al. (2013) Impact of the carbon and

nitrogen supply on relationships and connectivity between metabolism and

biomass in a broad panel of Arabidopsis accessions. Plant Physiol., 162,

347–363.

Sweetlove, L.J., Beard, K.F.M., Nunes-Nesi, A., Fernie, A.R. and Ratcliffe,

R.G. (2010) Not just a circle: flux modes in the plant TCA cycle. Trends

Plant Sci., 15, 462–470.

Sweetlove, L.J., Fell, D. and Fernie, A.R. (2008) Getting to grips with the plant

metabolic network. Biochem. J., 409, 27–41.

Sweetlove, L.J. and Ratcliffe, R.G. (2011) Flux-balance modeling of plant metab-

olism. Front. Plant Sci., 2, 38.

References

130

Sweetlove, L.J., Williams, T.C.R., Cheung, C.Y.M. and Ratcliffe, R.G. (2013)

Modelling metabolic CO₂ evolution--a fresh perspective on respiration.

Plant Cell Environ., 36, 1631–1640.

Taira, M., Valtersson, U., Burkhardt, B. and Ludwig, R.A. (2004) Arabidopsis

thaliana GLN2-encoded glutamine synthetase is dual targeted to leaf mito-

chondria and chloroplasts. Plant Cell, 16, 2048–2058.

Tcherkez, G., Cornic, G., Bligny, R., Gout, E. and Ghashghaie, J. (2005) In vivo

respiratory metabolism of illuminated leaves. Plant Physiol., 138, 1596–

1606.

Thiele, I. and Palsson, B.Ø. (2010) A protocol for generating a high-quality ge-

nome-scale metabolic reconstruction. Nat. Protoc., 5, 93–121.

Tomato Genome Consortium (2012) The tomato genome sequence provides in-

sights into fleshy fruit evolution. Nature, 485, 635–641.

Töpfer, N., Caldana, C., Grimbs, S., Willmitzer, L., Fernie, A.R. and Nikoloski,

Z. (2013) Integration of genome-scale modeling and transcript profiling re-

veals metabolic pathways underlying light and temperature acclimation in

Arabidopsis. Plant Cell, 25, 1197–1211.

Town, C.D., Cheung, F., Maiti, R., et al. (2006) Comparative genomics of Bras-

sica oleracea and Arabidopsis thaliana reveal gene loss, fragmentation, and

dispersal after polyploidy. Plant Cell, 18, 1348–1359.

Tschoep, H., Gibon, Y., Carillo, P., Armengaud, P., Szecowka, M., Nunes-Nesi,

A., Fernie, A.R., Koehl, K. and Stitt, M. (2009) Adjustment of growth

and central metabolism to a mild but sustained nitrogen-limitation in Ara-

bidopsis. Plant Cell Environ., 32, 300–318.

Valle, E.M., Boggio, S.B. and Heldt, H.W. (1998) Free Amino Acid Composition

of Phloem Sap and Growing Fruit of Lycopersicon esculentum. Plant Cell

Physiol., 39, 458–461.

Varma, A. and Palsson, B.O. (1995) Parametric sensitivity of stoichiometric flux

balance models applied to wild-type Escherichia coli metabolism. Biotech-

nol. Bioeng., 45, 69–79.

Vranová, E., Coman, D. and Gruissem, W. (2013) Network analysis of the MVA

and MEP pathways for isoprenoid synthesis. Annu. Rev. Plant Biol., 64,

665–700.

References

131

Vu, T.T., Stolyar, S.M., Pinchuk, G.E., et al. (2012) Genome-scale modeling of

light-driven reductant partitioning and carbon fluxes in diazotrophic unicel-

lular cyanobacterium Cyanothece sp. ATCC 51142. PLoS Comput. Biol., 8,

e1002460.

Walker, B.J., Strand, D.D., Kramer, D.M. and Cousins, A.B. (2014) The re-

sponse of cyclic electron flow around photosystem I to changes in

photorespiration and nitrate assimilation. Plant Physiol., 165, 453–462.

Wang, C., Guo, L., Li, Y. and Wang, Z. (2012) Systematic comparison of C3 and

C4 plants based on metabolic network analysis. BMC Syst. Biol., 6, 1–14.

Wang, H., Gu, M., Cui, J., Shi, K., Zhou, Y. and Yu, J. (2009) Effects of light

quality on CO2 assimilation, chlorophyll-fluorescence quenching, expres-

sion of Calvin cycle genes and carbohydrate accumulation in Cucumis

sativus. J. Photochem. Photobiol. B, 96, 30–37.

Wang, H., Jiang, Y.P., Yu, H.J., Xia, X.J., Shi, K., Zhou, Y.H. and Yu, J.Q. (2010) Light quality affects incidence of powdery mildew, expression of

defence-related genes and associated metabolism in cucumber plants. Eur.

J. Plant Pathol., 127, 125–135.

Wang, X., Heifetz, P.B., Gruissem, W. and Lange, M. (2003) Cross talk between

cytosolic and plastidial pathways of isoprenoid biosynthesis in Arabidopsis

thaliana. Proc. Natl Acad. Sci. U.S.A., 100, 6866–6871.

Wedig, C.L., Jaster, E.H. and Moore, K.J. (1987) Hemicellulose monosaccharide

compositionand in vitro disappearance of orchard grass and alfalfa hay. J.

Agr. Food Chem., 35, 214–218.

Weise, A., Barker, L., Kühn, C., Lalonde, S., Buschmann, H., Frommer, W.B.

and Ward, J.M. (2000) A new subfamily of sucrose transporters, SUT4,

with low affinity/high capacity localized in enucleate sieve elements of

plants. Plant Cell, 12, 1345–1355.

Wilhelm, C. and Selmar, D. (2011) Energy dissipation is an essential mechanism

to sustain the viability of plants: The physiological limits of improved pho-

tosynthesis. J. Plant Physiol., 168, 79–87.

Williams, T.C.R., Miguet, L., Masakapalli, S.K., Kruger, N.J., Sweetlove, L.J.

and Ratcliffe, R.G. (2008) Metabolic network fluxes in heterotrophic Ar-

abidopsis cells: stability of the flux distribution under different oxygenation

conditions. Plant Physiol., 148, 704–718.

References

132

Williams, T.C.R., Poolman, M.G., Howden, A.J.M., Schwarzlander, M., Fell,

D.A., Ratcliffe, R.G. and Sweetlove, L.J. (2010) A genome-scale meta-

bolic model accurately predicts fluxes in central carbon metabolism under

stress conditions. Plant Physiol., 154, 311–323.

Wingler, A., Lea, P.J., Quick, W.P. and Leegood, R.C. (2000) Photorespiration:

metabolic pathways and their role in stress protection. Philos. Trans. R. Soc.

Lond. Biol. Sci., 355, 1517–1529.

Wu, Q., Su, N., Shen, W. and Cui, J. (2014) Analyzing photosynthetic activity

and growth of Solanum lycopersicum seedlings exposed to different light

qualities. Acta Physiol. Plant., 36, 1411–1420.

Xiaoying, L., Shirong, G., Taotao, C., Zhigang, X. and Tezuka, T. (2014) Reg-

ulation of the growth and photosynthesis of cherry tomato seedlings by

different light irradiations of light emitting diodes (LED). Afr. J. Biotech-

nol., 11, 6169–6177.

XiaoYing, L., ShiRong, G., ZhiGang, X., XueLei, J. and Tezuka, T. (2011) Reg-

ulation of chloroplast ultrastructure, cross-section anatomy of leaves, and

morphology of stomata of cherry tomato by different light irradiations of

light-emitting diodes. HortScience, 46, 217–221.

Xiong, L. and Zhu, J.-K. (2003) Regulation of abscisic acid biosynthesis. Plant

Physiol., 133, 29–36.

Xiong, Y., DeFraia, C., Williams, D., Zhang, X. and Mou, Z. (2009) Deficiency

in a cytosolic ribose-5-phosphate isomerase causes chloroplast dysfunction,

late flowering and premature cell death in Arabidopsis. Physiol. Plant., 137,

249–263.

Xu, H., Zhang, J., Zeng, J., Jiang, L., Liu, E., Peng, C., He, Z. and Peng, X. (2009) Inducible antisense suppression of glycolate oxidase reveals its

strong regulation over photosynthesis in rice. J. Exp. Bot., 60, 1799–1809.

Yamamoto, H., Peng, L., Fukao, Y. and Shikanai, T. (2011) An Src homology 3

domain-like fold protein forms a ferredoxin binding site for the chloroplast

NADH dehydrogenase-like complex in Arabidopsis. Plant Cell, 23, 1480–

1493.

Yang, Y.-X., Wang, M.-M., Ren, Y., et al. (2014) Light-induced systemic re-

sistance in tomato plants against root-knot nematode Meloidogyne

incognita. Plant Growth Regul., 76, 167–175.

References

133

Yorio, N.C., Goins, G.D., Kagie, H.R., Wheeler, R.M. and Sager, J.C. (2001)

Improving spinach, radish, and lettuce growth under red light-emitting di-

odes (LEDs) with blue light supplementation. HortScience, 36, 380–383.

Yu, H. and Ong, B.-L. Effect of radiation quality on growth and photosynthesis of

Acacia mangium seedlings. Photosynthetica, 41, 349–355.

Yuan, H., Cheung, C.Y.M., Poolman, M.G., Hilbers, P.A.J. and Riel, N.A.W.

van (2016) A genome-scale metabolic network reconstruction of tomato

(Solanum lycopersicum L.) and its application to photorespiratory metabo-

lism. Plant J., 85, 289–304.

Zhang, X., Brammer, E., Pedersén, M. and Fei, X. (1997) Effects of light photon

flux density and spectral quality on photosynthesis and respiration in

Porphyra yezoensis (Bangiales, Rhodophyta). Phycol. Res., 45, 29–37.

135

Summary

Cellular metabolism is a complex process which encompasses all biochemical reac-

tions taking place within a cell. These biochemical reactions are organized into

numerous metabolic pathways which are often interconnected, shaping a dynamic

circuitry commonly referred as metabolic network. In living organisms, metabolic

network plays an important role in the cell survival. Studying the regulation of the

metabolic network will eventually contribute to understanding cellular mechanisms.

In this regard, mathematical and computational approaches, in particular, flux bal-

ance analysis (FBA), and related methods, are increasingly recognized as a powerful

tool to systematically study the organization of metabolic networks. In the present

thesis, we used the constraint-based modelling framework to investigate the func-

tions of plant biological processes and the effects of environmental perturbations,

such as availability of water and light.

Firstly, we examined how the modelling parameters commonly used in flux-

balanced analysis affect the flux distributions in plant metabolic networks (Chapter

2). The analyses were applied to Arabidopsis which is the best-investigated plant

model organism. Our analyses suggested that model structure has a large impact on

FBA predictions in comparison with the biomass composition when studying the

plant metabolic behavior using FBA. In addition, comparative analysis of the Ara-

bidopsis models paves the way for exploring the existence of principles that are

relevant for the regulation and robustness of plant central carbon metabolism.

Tomato (Solanum lycopersicum L.) is one of the most important vegetable

crops grown in the world. With the genome sequence available for tomato in 2012,

it is possible perform systems-level studies of this high-value vegetable crop. In this

thesis, we constructed a genome-scale metabolic model for tomato, iHY3410, which

is used for simulating tomato leaf metabolism (Chapter 3). The model is capable of

successfully describing the metabolic activities of well-known bio-chemical path-

ways for leaf cells under heterotrophic and phototrophic conditions.

Subsequently, we used the reconstructed tomato model iHY3410 to investigat-

ing the metabolic characteristics under drought stressed conditions (Chapter 4). The

model predictions suggested that the metabolic flux through the mevalonate (MVA)

pathway under drought conditions significantly differed from that under normal

conditions, which gave hints to possible adaptive metabolic responses to drought

Summary

136

stresses. The application of genome-scale networks to the abiotic stresses in plants

is still in its infancy, and we believe our tomato model can serve as a starting point

for studying the functional consequences of abiotic stresses.

Finally, the tomato model iHY3410 was extended to a light-specific model,

which was employed to explore the metabolic changes of tomato leaves in response

to changing light intensity and quality (Chapter 5). The comprehensive analyses in-

dicate mitochondrial respiration plays an important role in regulating the balance

between ATP and NADPH in plastid, thus enabling optimal growth performance in

varying light conditions. The interaction between the plastid and other organelles is

achieved by metabolite shuttles that differ with the light condition.

Additionally, this thesis discusses the potential extensions of the present work

(Chapter 6). Overall, the thesis work provides a systematic computational frame-

work for investigating metabolic activities under various growth conditions.

137

Publications

Yuan, H., Cheung, C.Y.M., Hilbers, P.A.J. and van Riel, N.A.W. Modelling the

effects of light intensity and quality on the light-driven metabolism of tomato. In

preparation.

Yuan, H., Cheung, C.Y.M., Hilbers, P.A.J. and van Riel, N.A.W. (2016) Flux bal-

ance analysis of plant metabolism: the effect of biomass composition and model

structure on model predictions. Front. Plant Sci.7, 537

Yuan, H., Cheung, C.Y.M., Poolman, M.G., Hilbers, P.A.J. and van Riel, N.A.W.

(2016) A genome-scale metabolic network reconstruction of tomato (Solanum lyco-

perscum L.) and its application to photorespiratory metabolism. Plant J., 85, 289–

304

Other publication outside the scope of this thesis

Golam Jalal Ahammed, Hui-Li Yuan, Joshua Otieno Ogweno, Yan-Hong Zhou,

Xiao-Jian Xia, Wei-Hua Mao, Kai Shi, Jing-Quan Yu (2016) Brassinosteroid alle-

viates phenanthrene and pyrene phytotoxicity by increasing detoxification activity

and photosynthesis in tomato. Chemosphere, 86, 546–555

138

Acknowledgements

Four years ago, I have never thought I would do a PhD abroad, nor switch my field

of study in experimental plant biology. But right now, I am completing my PhD in

computational biology at TU/e. I cannot believe it happened! This thesis would not

have been possible without the support and help from many people. I take this op-

portunity to express my gratitude to all the people who have helped and encouraged

me during my PhD.

First, I would like to thank my promoter prof.dr. Peter Hilbers for offering me

the opportunity to do my PhD study at Computational Biology group (CBio). Your

enthusiasm towards work has always impressed me and positively influenced me.

Both your broad knowledge and important insights greatly improved my work. Per-

haps, I was the most ignorant of my colleagues at the CBio group, but your trust and

encouragement kept me going during times I doubted it could be done.

I also would like to express my sincere gratitude to my daily supervisor prof.dr.

Natal van Riel, for his mentorship and support when I needed. I still remembered

the first conversations we had at Zhejiang University, China in 2012. When I told

you my worries for switching to a different filed, you said: I’m not telling you it’s

going to be easy, I’m telling you you can make it if you are really willing to learn.

Indeed, it was a challenging journey to me, which gets me frustrated and disap-

pointed with myself. But, your support and inspiration kept me going through this

tough journey. I truly appreciate the possibilities to attend conferences and to visit

other groups you provided, from which I have benefited a lot.

There is one person that contributed significantly to my PhD work and without

whom I would not be at this point, this person is Dr. Maurice Cheung from Yale-

NUS College, Singapore. Maurice, I was impressed by your rigor and enthusiasm

towards work. It was the ambition of being like you one day that motivated me and

gave me the strength to go beyond my limits. I have bothered you with so many

questions, but you were always patient with me. I cannot express more appreciation

for your guidance on my work, encouraging words in moments of self-doubt and

support whenever I needed. Without these, I would have had a much harder time. I

think I am fortunate enough to know you and work with you.

I also want to extend my gratitude to all my committee members: George

Ratcliffe, Bas Teusink, Jan Bergmans, Guofu Zhou and Maurice Cheung, thank all

Acknowledgements

139

of you for your time on reading my dissertation and valuable comments on my dis-

sertation. Dick de Ridder and Cees Oomens, thanks for serving on the public defense

of my dissertation.

I would like to acknowledge Dr. Mark Poolman from Oxford Brookes Univer-

sity, UK. Mark, thanks a lot for hosting me at Oxford Brookes University and for

the technique help on metabolic modelling in ScrumPy. I also want to thank all the

people in his group, especially, Hassan Hartman, Dipali Singh and Kailash Adhikari

for the instructions on ScrumPy and fruitful discussions. Dipali, thanks very much

for your help and caring during my stay in Oxford. It was always enjoyable to talk

with you and to do adventures with you. Kailash, thanks for sharing so many beau-

tiful songs with me, which I enjoyed a lot. I will never forget my wonderful time in

Oxford.

I also want to thank all my CBio group members, including the former mem-

bers, for sharing and for creating such a friendly atmosphere for work. Special

thanks to: Bram Wijnen, Joep Vanlier, Ceylan, Rik, Anne, Fianne, Yvonne, Zandra,

Sander, Ymke and Daniela for discussions on scientific or unscientific topics and

the PhD gatherings, from which I had a lot of fun, and of course, the nice foods.

Bram, I am very thankful for your valuable contribution to the project startup.

Thanks very much for picking me up from Schiphol airport and helping to settle

down when I moved to the Netherlands. More impressively, you prepared the Chi-

nese food for my first meals, which made me feel cared for. I also want to extend

my appreciation to your parents for their lovely hand-made Christmas cards, and for

the dinners we had together, which made me feel welcome.

Joep, many thanks for sharing your valuable experience and for giving helpful

suggestions for my PhD study. Doing a PhD, sometimes, doesn’t go smoothly, but

as you said, funfun

Ceylan, you like a sister for me, your caring gave me a lot of warmth, which

really alleviated me from feeling homesick, especially, in the first few months after

I came to the Netherlands. Many thanks for your inspirations when I was lost and

feeling down, and for sharing so many delicious desserts and candies with me.

Rik, you have a really good sense of humor. I had a lot of fun from your humor

and jokes. Thanks for inviting me for your spareribs, which I always miss a lot

Fianne, I am always impressed by your neat and appealing slides and/or posters

every time you presented. I think you are really skilled at that. You are always very

kindly and patient whenever I needed a favor. And it is always joyful to listen to

you while you talk about your Chinese courses.

Yvonne, thank you for inviting me to attend your wedding ceremony, wish you

and your husband a wonderful life together

Acknowledgements

140

Zandra, I always wonder how you can manage so many things in your life.

Your enthusiasm towards life and work is impressive. It was very great to do run-

nings with you, which were the cherished memories to me.

I also would like to acknowledge Rina, the secretary of our group for her ex-

cellent administrative support. Rina, your work really made our life much simpler.

Koen, thanks for all your effort in coordinating the laboratory. I highly appreciate

your help in improving the cover design. Dragan, thank you so much for all your

suggestions on my growth and your kindness to me.

My gratitude also goes to my friends Xuemin, Yuanyuan, Yu, Xuan and Wenya

(out of the Netherlands), Guiling, Xuefang, Zihan and all the people I met in Eind-

hoven. Thanks for all you have done for me throughout this journey. My thanks and

appreciation also go to Loka and Geremy from National University of Singapore for

their help and caring during my stay in Singapore. Thanks for bearing my craziness

and stubborn You guys made my stay in Singapore very joyful.

I acknowledge the Brainbridge project and China Scholarship Council (CSC)

for the financial support.

Finally, I would like to dedicate this thesis to my dearest parents for their sup-

port and love throughout all these years. I also want to thank my brothers, my sisters-

in-law, my sister and my brother-in-law for spoiling me, and for their support in all

my pursuits. I feel so lucky to have such a big family and to be your younger sister.

Completing a PhD is just the beginning, not the end. I am looking forward to

my next adventure.

Huili Yuan

October, 2016

Eindhoven, the Netherlands

141

Curriculum Vitae

Huili Yuan was born in Henan Province, China in 1986. She

received her Bachelor degree in Agronomy at Henan Agri-

cultural University, China in 2009. In the same year, she

started her master study in Horticulture in the group of

prof.dr. Jingquan Yu at Zhejiang University, China, and ob-

tained her Master degree in 2012. Her master thesis focused

on the effects of light quality on vegetable crops’ growth,

nutritional quality and disease resistance. From September

2012, she switched from discipline and started her PhD

study in Computational Biology group of prof.dr. Peter Hilbers at Eindhoven Uni-

versity of Technology (TU/e), the Netherlands, with a scholarship from China

Scholarship Council (CSC). She worked on the genome-scale metabolic modelling

of plant metabolism, under the supervision of prof.dr. Natal van Riel. During the

PhD study, she also collaborated with the research groups of dr. Mark Poolman at

Oxford Brookes University (UK) and dr. Maurice Cheung at Yale-NUS College

(Singapore) and visited these groups occasionally. The results of her PhD work are

presented in this dissertation.