171
The Poly-Cis Pathway of Carotene Desaturation: Enzymology, Herbicide Action and Retrograde Signaling Inaugural Dissertation to obtain the Doctoral Degree Faculty of Biology Albert-Ludwigs-Universität Freiburg im Breisgau Presented by Julian Koschmieder born in Bühl (Baden), Germany July 2017

The Poly-Cis Pathway of Carotene Desaturation: Enzymology

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

Page 1: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

The Poly-Cis Pathway of Carotene Desaturation:

Enzymology, Herbicide Action and Retrograde Signaling

Inaugural Dissertation to obtain the Doctoral Degree

Faculty of Biology

Albert-Ludwigs-Universität

Freiburg im Breisgau

Presented by

Julian Koschmieder

born in Bühl (Baden), Germany

July 2017

Page 2: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Dean of the Faculty of Biology: Prof. Dr. Bettina Warscheid

Promotion chairman: Prof. Dr. Andreas Hiltbrunner

Supervisor of work: Prof. Dr. Peter Beyer

1st Reviewer: Prof. Dr. Peter Beyer

2nd Reviewer: Prof. Dr. Ralf Reski

3rd Reviewer: Prof. Dr. Jens Timmer

Date of defense: 12th of September 2017

This work was carried out from April 2014 to July 2017 under the supervision of

Prof. Dr. Peter Beyer, Institute of Biology II, Cell Biology.

Page 3: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Major parts of this thesis have been published in:

Bruno M*, Koschmieder J*, Wüst F, Schaub P, Fehling-Kaschek M, Timmer J,

Beyer P, Al-Babili S (2016) Enzymatic study on AtCCD4 and AtCCD7 and their

potential to form acyclic regulatory metabolites. J Exp Bot 67: 5993–6005

* equal contribution

Brausemann A, Gemmecker S, Koschmieder J, Ghisla S, Beyer P, Einsle O

(2017) Structure of Rice (O. sativa) Phytoene Desaturase Provides Insights Into

Herbicide Binding and Reaction Mechanisms Involved In Carotene

Desaturation. Structure, in press, DOI: 10.1016/j.str.2017.06.002

Gemmecker S, Schaub P, Koschmieder J, Brausemann A, Drepper F,

Rodriguez-Franco M, Ghisla S, Warscheid B, Einsle O, Beyer P (2015) Phytoene

Desaturase from Oryza sativa: Oligomeric Assembly, Membrane Association

and Preliminary 3D-Analysis. PLoS One 10: e0131717

Schaub P, Wüst F, Koschmieder J, Yu Q, Virk P, Tohme J, Beyer P (2017) Non-

Enzymatic β-Carotene Degradation in (Provitamin A-Biofortified) Crop Plants.

J Agr Food Chem, in press

Submitted for publication:

Koschmieder J, Fehling-Kaschek M, Schaub P, Ghisla S, Brausemann A, Timmer

J, Beyer P (2017) Rice (O. sativa) Phytoene desaturase: a Functional

Evaluation of Structural Implications.

Wagner H, Beyer H, Engesser R, Hörner M, Ermes K, Koschmieder J, Beyer P,

Timmer J, Zurbriggen M, Weber W (2017) Synthetic Biology-Inspired Design of

Information Processing Materials Systems.

Page 4: The Poly-Cis Pathway of Carotene Desaturation: Enzymology
Page 5: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Table of Contents

I

Table of Contents

1 INTRODUCTION 1

1.1 Chemical structure of carotenoids 1

1.2 Carotenoids: biosynthesis and functions in plants 2

1.2.1 Functions in plants 2

1.2.2 General biosynthesis pathway in plants 3

1.2.3 Biosynthesis and function of carotenoid-derived

phytohormones 6

1.3 Poly-cis pathway of carotene desaturation in plants and

all-trans desaturation pathway in bacteria 8

1.3.1 Carotene desaturases in plants and bacteria 9

1.3.1.1 Plant-type phytoene desaturase PDS and ζ-carotene

desaturase ZDS 9

1.3.1.2 Bacterial phytoene desaturase CrtI 15

1.3.1.3 Bleaching herbicides targeting PDS 17

1.3.2 ζ-Carotene isomerase ZISO 22

1.4 Apocarotenoids as novel signaling compounds in plants 23

1.5 Objectives of this work 26

1.5.1 Biochemical characterization of PDS and ZDS 26

1.5.1.1 Kinetic and structural investigations on PDS and ZDS 26

1.5.1.2 Action mode of bleaching herbicides targeting PDS 27

1.5.2 Elucidation of the function of ζ-carotene isomerase ZISO 27

1.5.3 Identification of apocarotenoids originating from the poly-cis

pathway of carotene desaturation 28

2 MATERIALS 29

2.1 Chemicals and materials 29

2.2 Primers 31

2.3 Plasmids 32

2.4 Consumption materials 33

2.5 Bacterial strains 33

Page 6: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Table of Contents

II

3 METHODS 34

3.1 Nucleic acid methods 34

3.1.1 Isolation of nucleic acids 34

3.1.2 Separation of nucleic acids 34

3.1.3 Sequencing of nucleic acids 34

3.1.4 Polymerase chain reaction 34

3.1.5 Site-directed mutagenesis 34

3.1.6 Gene cloning 35

3.1.6.1 OsPDS-His6 and OsZDS-His6 35

3.1.6.2 (Mistic-)OsZISO-His6 35

3.1.7 Transformation of E. coli with plasmid DNA 36

3.2 Protein methods 36

3.2.1 Protein expression and purification 36

3.2.1.1 Expression and IMAC purification of OsPDS-His6 36

3.2.1.2 Expression and IMAC purification of OsZDS-His6 36

3.2.1.3 Expression and crude preparation of Mistic-OsZISO-His6 37

3.2.1.4 Expression and IMAC purification of LeCRTISO-His6 37

3.2.1.5 Expression and crude preparation for AtCCDs 38

3.2.1.6 Gel permeation chromatography for OsPDS-His6 38

3.2.1.7 Gel permeation chromatography for OsZDS-His6 38

3.2.2 Enzymatic assays 39

3.2.2.1 OsPDS-His6 in vitro 39

3.2.2.2 OsZDS-His6 in vitro 39

3.2.2.3 Mistic-OsZISO-His6 in vitro 39

3.2.2.4 LeCRTISO-His6 in vitro 39

3.2.2.5 AtCCDs in vitro 40

3.2.2.6 OsZISO-His6 in vivo 40

3.2.3 SDS-PAGE 40

3.2.4 Protein precipitation 41

3.2.4.1 Ammonium sulfate precipitation 41

3.2.4.2 Chloroform-methanol precipitation 41

Page 7: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Table of Contents

III

3.2.5 Protein quantification 41

3.2.5.1 Bradford assay 41

3.2.5.2 Nanodrop 42

3.2.6 Membrane association assay 42

3.2.7 Preparation of carotenoid-containing liposomes 42

3.2.8 Photometric quantification of protein-bound FAD 42

3.3 Chromatography 42

3.3.1 High performance liquid chromatography 42

3.3.2 Extraction and purification of carotenoids from

tangerine tomato and carotenoid-containing bacteria 44

3.3.3 Effective liposomal concentrations of carotenes,

decylplastoquinone and norflurazon 45

3.3.4 Liquid chromatography – mass spectrometry

3.3.4.1 Identification of apocarotenoids and carotene

deuteration analysis 45

3.3.4.2 Identification of C35 carotenoids 45

3.3.4.3 Identification of protein-bound nucleotide cofactors 46

3.3.5 Gas chromatography – mass spectrometry and

solid phase micro extraction 46

3.4 Cryo scanning electron microscopy 46

3.5 Homology modeling and in silico docking of OsZDS-His6 47

3.6 Mathematical modeling of PDS reaction time courses

and kinetics 47

3.7 Deduction of non-covalent interactions from changes in

electron density gradients 48

4 RESULTS 49

4.1 Biochemical characterization of phytoene desaturase PDS 49

4.1.1 Association and oligomerization of OsPDS-His6 at

liposomal surfaces 49

4.1.1.1 Monotopic association of OsPDS-His6 with liposomes 49

Page 8: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Table of Contents

IV

4.1.1.2 Homooligomerization of OsPDS-His6 at liposomal

membrane surfaces 50

4.1.2 Regio-specificity of carotene desaturation by OsPDS-His6 51

4.1.2.1 Regio-specificity with 15-cis-penta-nor-phytoene (C35) 53

4.1.2.2 Regio-specificity upon mutation of the substrate

cavity back end 55

4.1.3 Reaction mechanism of carotene desaturation in OsPDS-His6 56

4.1.4 Kinetic mechanism of the bi-substrate reaction in OsPDS-His6 59

4.2 Kinetic characterization of PDS and mathematical modeling 61

4.2.1 Basic characterization of the OsPDS-His6 reaction 61

4.2.2 Mathematical modeling of OsPDS-His6 reaction time courses 64

4.2.3 Substrate concentration-dependent kinetics of OsPDS-His6 76

4.3 Investigations on herbicide resistance in PDS 80

4.3.1 Kinetic analysis of the mode of inhibition by

norflurazon for OsPDS-His6 80

4.3.2 Norflurazon resistance conferred by point mutations 82

4.3.3 Identification of non-covalent norflurazon – OsPDS-His6

interactions 87

4.4 ζ-Carotene desaturase ZDS 89

4.4.1 Basic biochemical characterization of OsZDS-His6 89

4.4.1.1 Establishing native OsZDS-His6 purification and

an in vitro assay 89

4.4.1.2 Identification of nucleotide redox cofactors bound

to OsZDS-His6 91

4.4.1.3 Kinetic mechanism of the OsZDS-His6 bi-substrate reaction 92

4.4.2 Structure and reaction mechanism of OsZDS-His6 93

4.4.2.1 Reaction mechanism of carotene desaturation

by OsZDS-His6 93

4.4.2.2 Homology modeling of OsZDS-His6 and analysis of

the active site 94

4.4.2.3 Homooligomerization of OsZDS-His6 in solution 97

4.4.3 Kinetic characterization of OsZDS-His6 99

Page 9: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Table of Contents

V

4.4.3.1 Basic characterization and optimization of

reaction parameters 99

4.4.3.2 Reaction time course of ζ-carotene conversion

by OsZDS-His6 100

4.4.3.3 Substrate concentration dependency of the

OsZDS-His6 reaction 101

4.5 ζ-Carotene isomerase ZISO 103

4.5.1 Substrate specificity of OsZISO-His6 in E. coli 103

4.5.2 Reconstitution of the poly-cis pathway of

carotene desaturation 105

4.5.3 Isomerization of 9,15,9’-tri-cis-ζ-carotene by ZISO in vitro 106

4.6 Apocarotenoids derived from the poly-cis pathway

of carotene desaturation 108

4.6.1 AtCCD4-mediated cleavage of acyclic

poly-cis carotenes 109

4.6.2 AtCCD7-mediated cleavage of acyclic

poly-cis carotenes 112

4.6.3 Identification of apocarotenoids originating from

acyclic 9-cis-carotenes by AtCCD7 cleavage 115

4.6.4 Secondary cleavage of 9-cis-ζ-apo-10’-carotenal

by AtCCDs 119

5 DISCUSSION 122

5.1 Phytoene desaturase PDS: evaluation of implications

deduced from the OsPDS-His6 structure 122

5.1.1 OsPDS-His6 is a monotopic membrane protein 122

5.1.2 OsPDS-His6 employs an unprecedented “flavin only”

reaction mechanism during carotene desaturation 123

5.1.3 OsPDS-His6 follows an ordered ping pong

bi bi mechanism 123

5.1.4 Substrate and regio-specificity of PDS: role of

the 15-cis-configuration 125

Page 10: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Table of Contents

VI

5.1.5 Kinetic characterization and mathematical

modeling of OsPDS-His6 128

5.1.5.1 Modeling supports substrate channeling in

OsPDS-His6 homotetramers 128

5.1.5.2 Rate-limiting steps in OsPDS-His6 130

5.1.6 Characterization of norflurazon - OsPDS-His6 interactions 132

5.1.6.1 Quinone-competitive inhibition by norflurazon is

mediated by several non-covalent interactions 132

5.1.6.2 Site-directed mutagenesis confers norflurazon

resistance at the expense of catalytic activity 135

5.2 ζ-carotene desaturase ZDS: a comparison to PDS 138

5.2.1 OsZDS-His6 utilizes an ordered ping pong bi bi

mechanism and a “flavin only” mechanism 138

5.2.2 OsZDS-His6 might not employ homooligomerization

and substrate channeling 139

5.3 ζ-Carotene Isomerase ZISO: a bona fide enzyme with

the potential for additional functions 141

5.4 The poly-cis pathway of carotene desaturation

and retrograde signaling 143

5.4.1 AtCCD7 forms linear 9-cis-apocarotenoids as

potential signaling molecule precursors 143

5.4.2 Linear 9-cis-apocarotenoids are not converted into

strigolactone-like metabolites by AtCCD8 145

6 SUMMARY 147

7 REFERENCES 149

8 ACKNOWLEDGEMENTS 157

Page 11: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Abbreviations

VII

Abbreviations

A. thaliana Arabidopsis thaliana

ABA abscisic acid

Acc. no. accession number

APCI atmospheric pressure chemical ionization

BCH β-carotene hydroxylase

bp base pairs

CCD carotenoid cleavage dioxygenase

CCO carotenoid cleavage oxygenase

ceff effective liposomal concentration

CHAPS 3-((3-cholamidopropyl)dimethylammonio)-1-

propanesulfonate

Chl chlorophyll

CMC critical micellar concentration

CrtI bacterial phytoene desaturase

CRTISO carotene isomerase

CYP cytochrome P450 hydroxylase

DFT density functional theory

DMAPP dimethylallyl diphosphate

DNA desoxyribonucleic acid

dNTP desoxyribonucleotide triphosphate

DPQ decylplastoquinone

E. coli Escherichia coli

e.g. for example (exempli gratia)

EDTA ethylenediaminetetraacetic acid

FAD flavin adenine dinucleotide

FMN flavin adenine mononucleotide

GC-MS gas chromatography mass spectrometry

GGPP geranylgeranyl diphosphate

GGPPS geranylgeranyl diphosphate synthase

GPC gel permeation chromatography

Page 12: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Abbreviations

VIII

GST glutathione S-transferase

HPLC high performance liquid chromatography

i.e. that is (id est)

IMAC immobilized metal affinity chromatography

IPP isopentenyl diphosphate

IPTG isopropyl-β-D-thiogalactopyranoside

kDa kilodalton

LC-MS liquid chromatography mass spectrometry

LHC light harvesting complex

M molar

mAU milli arbitrary unit

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

MES 2-(N-morpholino)ethanesulfonic acid

min minutes

n.d. not determined

NAD nicotinamide adenine dinucleotide

NADP nicotinamide adenine dinucleotide phosphate

NCED nine-cis-epoxycarotenoid dioxygenase

NDH NAD(P)H:quinone oxidoreductase

NFZ norflurazon

nm nanometer

NPQ non-photochemical quenching

NXS neoxanthin synthase

O. sativa Oryza sativa

ODE ordinary differential equation

PAGE poly acrylamide gelectrophoresis

PCR polymerase chain reaction

PDS phytoene desaturase

PQ plastoquinone

PSY phytoene synthase

PTOX plastid terminal oxidase

Page 13: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Abbreviations

IX

Q quinone

ROS reactive oxygen species

SDS sodium dodecyl sulfate

SEM scanning electron microscopy

SL strigolactone

TBME tert-butylmethylether

TCEP tris(2-carboxyethyl)phosphine

TEMED N,N,N’,N’-tetramethylethylendiamin

TLC thin layer chromatography

Tris tris(hydroxymethyl)aminomethane

UTR untranslated region

v/v volume per volume

VDE violaxanthin deepoxidase

w/v weight per volume

wt wild type

ZDS ζ-carotene desaturase

ZEP zeaxanthin epoxidase

ZISO ζ-carotene Isomerase

β-LCY lycopene-β-cyclase

ε-LCY lycopene-ε-cyclase

Page 14: The Poly-Cis Pathway of Carotene Desaturation: Enzymology
Page 15: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

1

1 Introduction

1.1 Chemical structure of carotenoids

Carotenoids comprise a group of more than 700 isoprenoids that are mostly

C40 compounds (Britton et al., 2004). They originate from the condensation of

six units of isopentenyl diphosphate (IPP) and two units of its isomer,

dimethylallyl diphosphate (DMAPP). Carotenoids can be either acyclic

hydrocarbons or have cyclic ionone end groups and are subdivided into non-

oxygenated carotenes and oxygenated xanthophylls. All carotenoids possess

a polyene system consisting of 3 to 15 conjugated double bonds and

absorbing light between 285 nm and 500 nm. Depending on the number of

double bonds present, this chromophore confers a yellow to red color. Albeit

trivial names are commonly used, C40 carotenoids are systematically

numbered according to IUPAC (International Union of Pure and Applied

Chemistry) guidelines to unambiguously assign positions, such as of cis or trans

double bonds, cleavage sites or substituents. Numbering for carotenoids starts

from C1 or C1’ at the outermost carbon of the backbone and is proceeded

towards the central C15 or C15’ at the molecule symmetry center (Fig.1-1 A).

Counting is continued from the outermost methyl position C16 or C16’

onwards to methyl groups C20 or C20’.

Fig.1-1 Structure of carotenoids and numbering of carbon atoms.

(A) Structure and numbering of all-trans-ζ-carotene as an acyclic carotene. (B)

Structure and numbering of all-trans-ζ-apo-10-carotenal. (C) Structure and carbon

numbering in cyclic carotenoids and apocarotenoids.

A

B CR

Page 16: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

2

Apocarotenoids represent oxidative cleavage products of carotenoids. They

are named after their carotenoid precursor, the carbon position adjacent to

the cleaved double bond and the oxygen function they contain (suffixes: -al,

aldehyde; -one, ketone). For instance, ζ-apo-10-carotenal originates from

cleavage of ζ-carotene next to C9-C10 and possesses an aldehyde function

(Fig.1-1 B). The structure and numbering of carbons in cyclic carotenoids is

given in Fig.1-1 C.

1.2 Carotenoids: biosynthesis and functions in plants

1.2.1 Functions in plants

In plants, carotenoids exert essential photosynthetic functions which are

based on their physicochemical properties. Carotenoids, xanthophylls in

particular, are major pigments in light harvesting complexes (LHC). They

expand the absorption spectrum of LHC chlorophylls (Chl) and contribute to

photosynthetic light reactions by transferring excitation energy from the

antenna towards reaction centers via Förster resonance energy transfer (for

review, see Demmig-Adams et al., 1996).

Furthermore, they play a key role in protecting plant cells from photooxidative

damage in two different ways. In contrast to Chl, they easily dissipate

excitation energy thermally after excitation, a process called non-

photochemical quenching (NPQ) (for review, see Jahns and Holzwarth, 2012).

They consequently diminish the formation of 3Chl and subsequently of 1O2

and other reactive oxygen species (ROS) that would otherwise damage cell

components. Furthermore, the xanthophyll cycle – in which zeaxanthin,

antheraxanthin and violaxanthin are rapidly interconverted by ZEP and VDE

(see 1.2.2) – is essential to adaptation of NPQ capacity according to light

supply. The existence of a second xanthophyll cycle consisting of lutein and

lutein-epoxide remains controversial. The photoprotective essentiality of

carotenoids is underscored by the mode of action of bleaching herbicides –

e.g. norflurazon, diflufenican, fluridone or J852. They are carotenoid

Page 17: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

3

biosynthesis inhibitors whose application causes lethal photooxidative

damage. Moreover, carotenoids are antioxidants and quench ROS by their

own oxidation which frequently leads to their cleavage. Additionally, β-

cyclocitral, a cleavage product of β-carotene, serves as a mediator of

adaptation to light stress by inducing genes involved in ROS scavenging (for

review, see Ramel et al., 2012 and Havaux et al., 2014).

Moreover, carotenoids exert structural functions by stabilizing and organizing

components of plastid membranes and photosynthetic complexes (Havaux,

1998; Gruszecki and Strzalka, 2005). Plastid membrane fluidity is strongly

influenced by xanthophyll cycle carotenoids. Zeaxanthin is presumably

released into the membrane during light stress, converted to violaxanthin with

a higher NPQ capacity and rebound to antenna complexes. Both

xanthophylls lower membrane fluidity and improve thermal membrane

stability in their transiently unbound state (Havaux et al., 1998). Furthermore,

carotenoids are essential for the functional assembly, structural integrity and

stability of light harvesting complexes, photosystem II and of the Cytb6f

complex (Plumley and Schmidt, 1987; Santabarbara et al., 2013). Finally,

carotenoids serve as colorants in tissues such as fruits or flowers and contribute

to the attraction of pollinators and the zoochoric dispersal of seeds (for

review, see Cazzonelli, 2011).

1.2.2 General biosynthesis pathway in plants

In plants, carotenoid biosynthesis takes place in plastids catalyzed by

membrane-associated enzymes (for a review on subplastid localization, see

Joyard et al., 2009). The plant-type biosynthetic pathway (Fig.1-2) is of

cyanobacterial origin and is regulated at multiple levels (for review, see

Cazzonelli und Pogson, 2010). Carotenoid biosynthesis starts from the central

metabolites isopentenyl diphosphate (IPP) and its isomer dimethylallyl

diphosphate (DMAPP). In plants, these precursors are not provided by the

cytosolic mevalonate pathway but by the methylerythritol-4-phosphate

pathway in the plastid stroma (for review, see Rodriguez-Concepción and

Page 18: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

4

Boronat, 2002). Geranylgeranyl diphosphate synthase (GGPPS) yields the

central C20 metabolite GGPP – a precursor involved in the formation of

several plant metabolites, e.g. chlorophylls, tocopherols or gibberellins.

GGPPS is a prenyltransferase condensing one DMAPP and three IPP in a head

to tail fashion (Dogbo et al., 1988). Subsequently, phytoene synthase (PSY)

catalyzes the first committed step in carotenoid biosynthesis by condensing

two molecules of GGPP to yield 15-cis-phytoene (Dogbo et al., 1988). Due to

the hydrophobicity of phytoene (and all carotenoids), PSY is thought to be

attached to thylakoids when active in order to deposit phytoene into the

hydrophobic membrane core (Schledz et al., 1996; Welsch et al., 2000).

Phytoene formation is followed by the so-called poly-cis pathway of

carotenoid desaturation (Fig.1-2) in which 15-cis-phytoene is desaturated and

isomerized to all-trans-lycopene by two desaturases and two cis-trans

isomerases. The poly-cis configuration of the carotenes involved was

elucidated by nuclear magnetic resonance spectroscopy (Clough and

Pattenden, 1979). The desaturation reactions extend the chromophore from a

triene in phytoene to an undecaene in lycopene resulting in a maximal

absorption at 470 nm and a red color. Phytoene desaturase (PDS)

symmetrically desaturates 15-cis-phytoene twice and yields the intermediate

9,15-di-cis-phytofluene and the end product 9,15,9’-tri-cis-ζ-carotene (Bartley

et al., 1999). ζ-carotene-isomerase (ZISO) isomerizes 9,15,9’-tri-cis-ζ-carotene

to 9,9’-di-cis-ζ-carotene (Li et al., 2007; Chen et al., 2010) which is then

symmetrically desaturated twice by ζ-carotene desaturase (ZDS) to yield the

intermediate 7,9,9’-tri-cis-neurosporene (syn. proneurosporene) and end

product 7,9,7’,9’-tetra-cis-lycopene (syn. prolycopene; Beyer et al., 1989;

Breitenbach and Sandmann, 2005). Finally, carotene isomerase (CRTISO)

symmetrically isomerizes prolycopene twice to all-trans-lycopene via the

intermediate 7,9-di-cis-lycopene (Isaacson et al., 2002; Isaacson et al., 2004;

Yu et al., 2011). The poly-cis pathway of carotene desaturation is one of the

major topics of this work and will be dealt with in greater detail in 1.3.1.

Page 19: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

5

Fig. 1-2 Biosynthesis of carotenoids and carotenoid-derived phytohormones in plants.

The biosynthetic pathway of carotenoids, including the poly-cis pathway of

carotene desaturation, is shown. Links to strigolactones and abscisic acid

biosynthesis are highlighted. The all-trans pathway of carotene desaturation in

bacteria and fungi, mediated by phytoene desaturase CrtI, is indicated by a red

arrow. GGPP, geranylgeranyl diphosphate; PSY, phytoene synthase; PDS, phytoene

desaturase; ZISO, ζ-carotene isomerase; ZDS, ζ-carotene desaturase; CRTISO,

carotene isomerase; β-LCY, lycopene-β-cyclase; ε-LCY, lycopene-ε-cyclase; BCH, β-

carotene hydroxylase; CYP97A/B, cytochrome P450 hydroxylases; ZEP, zeaxanthin

epoxidase; VDE, violaxanthin deepoxidase NXS, neoxanthin synthase.

15-cis-phytoene

9,15,9‘- tri-cis-ζ-carotene

9,9‘-di-cis-ζ-carotene

7,9,7‘,9‘-tetra-cis-lycopene

all-trans-lycopene

GGPP2x

β-carotene

zeaxanthin

antheraxanthin

violaxanthin

neoxanthin

α-carotene

lutein

PSY

PDS

ZDS

CRTISO

ZISO

NXS

ZEP

BCH

β-LCYβ-LCYε-LCY

CYP97A CYP97C

abscisic acid

strigolactones

poly-cis pathway

VDE

ZEPVDE

CrtI

Page 20: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

6

Subsequent to desaturation, all-trans-lycopene is cyclized by lycopene-β-

cyclase (β-LCY) and / or lycopene-ε-cyclase (ε-LCY) introducing β-ionone or

ε-ionone rings, respectively, which only differ in the position of one double

bond (C5=C6 in β-ionone, C4=C5 in ε-ionone; Yu et al., 2010; Yu and Beyer,

2012; Cunningham et al., 1996). The introduction of two β-ionone rings at the

ends of the lycopene backbone yields β-carotene, whereas introduction of

one β- and ε-ionone ring results in α-carotene.

The final phase of the pathway includes various oxygenation reactions of α-

carotene and β-carotene, resulting in xanthophyll formation. Hydroxylases

with ε-/β-ring specificities hydroxylate at C3 / C3’ (for review, see Tian et al.,

2004 and Kim et al., 2009): α-carotene is first converted to zeinoxanthin by a ε-

hydroxylase and then to lutein by a β-hydroxylase. β-Carotene is hydroxylated

twice by a β-hydroxylase to yield zeaxanthin. Zeaxanthin epoxidase (ZEP)

introduces epoxy functionalities to form antheraxanthin and violaxanthin

(Bouvier et al., 1996). In the context of the so-called xanthophyll cycle (see

1.2.1), these reactions are reversed by violaxanthin deepoxidase (VDE) (Bugos

et al., 1998; Arnoux et al., 2009; Fufezan et al., 2012). It remains unclear how

neoxanthin biosynthesis is accomplished. The involvement of a specific

neoxanthin synthase (NXS) or of a bifunctional β-LCY enzyme is debated

(North et al., 2007; Al-Babili et al., 2000; Bouvier et al., 2000). Furthermore, it

remains elusive whether or not a 9-cis-isomerase for violaxanthin and

neoxanthin exists in order to provide the precursors for abscisic acid

biosynthesis (North et al., 2007; see 1.2.3).

1.2.3 Biosynthesis and function of carotenoid-derived phytohormones

Two important phytohormones are biosynthesized from carotenoids through

oxidative cleavage catalyzed by carotenoid cleavage oxygenases (CCOs):

abscisic acid (ABA) and strigolactones. CCOs use molecular oxygen to

cleave double bonds in their carotenoid substrates, yielding two

apocarotenoid products with aldehyde or keto functionalities.

Page 21: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

7

ABA biosynthesis (for review, see Nambara and Marion-Poll, 2005) is initiated

by the oxidative cleavage of 9-cis-violaxanthin and 9-cis-neoxanthin at the

C11-C12 double bond by a 9-cis-epoxy-carotenoid cleavage dioxygenase

(NCED) – and yields the C15 ABA precursor xanthoxin. In the cytosol, xanthoxin

is then converted into ABA via abscisic aldehyde. ABA is a phytohormone

with versatile functions (for review, see Nambara and Marion-Poll, 2005), e.g.

induction of abiotic and biotic stress tolerance, regulation of stomata closure

and water balance, seasonal leaf abscission as well as the control of seed

dormancy in an interplay with gibberellic acid.

Strigolactones (SL) comprise a structurally and functionally diverse group of

phytohormones. However, they all share a core biosynthetic pathway (for

review, see Ruyter-Spira et al., 2013 and Al-Babili and Bouwmeester, 2015). SLs

derive from β-carotene which is first isomerized to 9-cis-β-carotene by the

isomerase D27 and then cleaved by CCD7 to form 9-cis-β-apo-10’-carotenal.

The latter is converted to the universal strigolactone precursor carlactone by

CCD8 by simultaneously occurring oxidative cleavage and complex

intramolecular rearrangement reactions (Alder et al., 2012; Bruno et al., 2017).

Carlactone is then further modified by P450 enzymes and hydroxylases to form

the sensu stricto SLs, such as 5’-deoxystrigol. SLs shape plant architecture by

enforcing apical dominance, i.e. inhibition of shoot branching, and by

regulating lateral root formation and root hair growth in response to nutrient

deprivation. Furthermore, SLs are involved in symbiotic as well as parasitic

relations after their secretion into the soil. On the one hand, SLs facilitate the

establishment of a symbiosis with arbuscular mycorrhizal fungi and serve as

inducers of hyphal branching. On the other hand, SLs are exploited by

parasitic plants of the Striga und Orobanche genera. Their seeds only

germinate in the presence of SLs that indicate the presence of a host plant

on which they are obligatorily dependent.

Page 22: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

8

1.3 Poly-cis pathway of carotene desaturation in plants and all-trans

desaturation pathway in bacteria

In the 1970s cis-carotenes were found as biosynthetic intermediates in fruits of

the tangerine mutant of tomato (Qureshi et al., 1974). Their stereo-

configuration was assigned by nuclear magnetic resonance (NMR)

spectroscopy, identifying 15-cis-phytoene, 9,15-di-cis-phytofluene, 9,9’-di-cis-

ζ-carotene, 7,9,9’-tri-cis-neurosporene and 7,9,7’,9’-tetra-cis-lycopene

(Clough and Pattenden, 1979). These isomers were also found in the mutant

CD-6 of the green algae Scenedesmus obliquus (Ernst and Sandmann, 1988).

With the identification of carotene isomerase (CRTISO; Isaacson et al., 2002;

Park et al., 2002), being affected in the above mentioned mutants, and with

the demonstration that the wild type carotenogenic enzymes of A. thaliana

yield poly-cis carotenes (Bartley et al., 1999), the ubiquitous occurrence of the

poly-cis carotene desaturation pathway in plants became generally

accepted. In contrast, the bacterial phytoene desaturase products were

trans-configured (Linden et al., 1991; Fraser et al., 1992). As of today, plant

carotene desaturation employs two homologous desaturases PDS and ZDS

and the two cis-trans isomerases ZISO and CRTISO (see 1.2), together yielding

all-trans-lycopene from 15-cis-phytoene. The all-trans pathway of carotene

desaturation in bacterial carotenogenesis employs only one enzyme, namely

phytoene desaturase CrtI, to produce the identical product from the

identical precursor (Fig.1-2, red arrow). It remains enigmatic what favored the

evolution of the complex poly-cis pathway in plants. One hypothesis is that its

complexity allows better fine-tuning of pathway flux (see 1.3.1.1; Sandmann,

2009). Recent findings suggest that apocarotenoids originating from the poly-

cis pathway are involved in plastid development and biosynthetic feedback

regulation (see 1.4).

PDS, ZDS and ZISO are the subjects of this work. The current knowledge of

these enzymes is summarized below.

Page 23: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

9

1.3.1 Carotene desaturases in plants and bacteria

1.3.1.1 Plant-type phytoene desaturase PDS and ζ-carotene desaturase ZDS

Phytoene desaturase (PDS) has a long history of research but remains poorly

understood. PDS and its cyanobacterial ortholog CrtP convert 15-cis-

phytoene via two formally identical reactions at both half sides of the

symmetrical substrate. Trans-configured double bonds are introduced at C11-

C12 and C11’-C12’ and concomitantly, trans-to-cis isomerization occurs at

the existing double bond at C9-C10 and C9’-C10’. Consequently, PDS yields

the intermediate 9,15-di-cis-phytofluene and the end product 9,15,9’-tri-cis-ζ-

carotene (Fig. 1-3 A; Bartley et al., 1999; Breitenbach and Sandmann, 2005;

for stereo-configuration, see Clough and Pattenden, 1979).

Fig.1-3 Scheme of the PDS reaction with the involved carotene stereoisomers and

homotetramerization of OsPDS-His6 in crystallo.

(A) 15-cis-phytoene is converted by PDS to yield 9,15,9’-tri-cis-ζ-carotene via the

intermediate 9,15-di-cis-phytofluene. Desaturation sites are highlighted in red. The

9,15-di-cis-phytofluene

15-cis-phytoene

9,15,9‘-tri-cis-ζ-carotene

FADoxPQ

PQH2

hνNDH

PTOX

2e-

FADred

FADoxPQ

PQH2

hνNDH

PTOX

2e-

FADred

12

11

12‘

11‘

109

9‘10‘

substrate cavity entranceA B

Page 24: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

10

enzyme-bound redox reactions involving FAD and plastoquinone as well as redox

chains that influence the PQ/PQH2 ratio - and consequently PDS activity - are

depicted. PQ, plastoquinone; PQH2, plastohydroquinone; hν, light energy

(photosynthetic electron transport chain); NDH, NAD(P)H:quinone oxidoreductase;

PTOX, plastid terminal oxidase. (B) Structure of OsPDS-His6 homotetramer obtained

upon protein crystallography (modified from Brausemann et al., 2017). Every

monomer is given in a different color and the substrate cavities are indicated.

Sequence alignments revealed that PDS possesses a (di)nucleotide cofactor

binding motif at the N-terminus, the Rossmann fold (for review, see Hanukoglu,

2015), and a stretch of 22 conserved hydrophobic amino acids has been

suggested to represent the phytoene binding site (Pecker et al., 1992). PDS

was shown to bind flavin adenine dinucleotide (FAD) as a redox cofactor (Al-

Babili et al., 1996; Hugueney et al., 1992). Al-Babili et al. (1996) observed that

PDS is targeted to plastids by a cleavable N-terminal transit peptide and

localized in the stroma in association with the chaperon Hsp70 as an

unflavinylated and inactive species. This is in line with the findings by Bonk et

al. (1997) which are related to the fate of the immature protein after in vitro

import into pea chloroplasts. A second PDS species is bound to plastid

membranes, flavinylated and enzymatically active. A concerted flavinylation

and membrane association of PDS assisted by chaperones was suggested

(Al-Babili et al., 1996) because membrane association in the absence of FAD

led to inactive PDS that was unable to bind this cofactor. Consequently, PDS

binds FAD as a non-exchangeable redox cofactor requiring its reoxidation

after every redox reaction for repeated catalysis (Al-Babili et al., 1996).

In the past, redox processes during PDS desaturation have been investigated

in complex systems. In daffodil chromoplasts, PDS and ZDS were only active in

the presence of oxygen that was replaceable by oxidized quinones (Beyer et

al., 1989, Mayer et al., 1990). Consequently, quinones were considered as

exchangeable electron acceptors requiring subsequent reoxidation by

oxygen (Mayer et al., 1990). In accordance with this, findings of Norris et al.

(1995) showed that PDS was inactive in the plastoquinone(PQ)–deficient

Page 25: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

11

Arabidopsis mutants pds1 and pds2. First in vitro experiments with

heterologously expressed Gentiana lutea PDS and Synechococcus CrtP

supported this notion (Breitenbach et al., 2001). Thus, FADox acts as an

intermediary electron acceptor. After reduction by electrons originating from

carotene desaturation, FAD reoxidation by PQ as an exchangeable terminal

electron acceptor is required to regenerate FADox for another round of

catalysis.

PDS activity strongly depends on the redox state of the plastid PQ pool that is

finely regulated by a network of redox chains (for review, see Mc Donald et

al., 2011 and Carol and Kuntz, 2001; Fig.1-3; Fig.1-4). In green tissues, it

primarily depends on the relative activities of the Cytb6f complex and the

photosystems, i.e. the photosynthetic electron transport. In non-

photosynthetic tissues, two additional redox systems come into play. First,

plastid terminal oxidase PTOX mediates plastohydroquinone (PQH2)

reoxidation by oxygen (Josse et al., 2003), explaining the dependence of PDS

activity on oxygen observed in daffodil chromoplasts (Beyer et al., 1989;

Mayer et al., 1990). In chromoplasts, PTOX is essential for carotenogenesis as it

provides plastoquinone whereas it is dispensable in photosynthetic tissues

when chloroplasts and photosynthetic electron transport are fully developed

(Carol and Kuntz, 2001). Second, a NAD(P)H:quinone oxidoreductase was

described in chromoplasts that mediates quinone reduction with NAD(P)H,

providing plastohydroquinone (Nievelstein et al., 1995). Interestingly, a

PQ/PQH2 redox optimum is required in chromoplasts to attain optimal PDS

activity (Nievelstein et al., 1995). This might suggest that PQ does not simply

act as a PDS electron acceptor. Thus, PDS has been envisioned as being

integrated into the above described redox chains that cover the redox

potential difference of the PDS-bound FAD/FADH2 and O2/H2O redox pairs of

approximately E0’ = 0.7 – 0.9 V. It remains to be elucidated whether the

accompanying ∆G0 is thermodynamically required to drive the PDS reaction,

i.e. whether carotene desaturation needs to be energetically coupled to PQ

and oxygen reduction. Otherwise, the participation of PQ and oxygen would

affect the PDS reaction only kinetically.

Page 26: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

12

Fig.1-4 Plastoquinone-mediated redox control of the PDS and ZDS reactions.

PDS and ZDS utilize plastoquinones as diffusible terminal electron acceptor to

reoxidize their bound FAD after carotene desaturation. The redox state of the

plastoquinone pool in plastid membranes controls the plastoquinone supply and

therefore, PDS activity. Thus, PDS activity is influenced by three redox chains that

utilize plastoquinones as electron donor or acceptor, namely PTOX and NDH (mainly

relevant in chromoplasts and non-photosynthetic tissues) as well as the

photosynthetic electron transport chain with PSII and the Ctyb6f complex. NDH,

NAD(P)H:quinone oxidoreductase; PTOX, plastid terminal oxidase, PSII, photosystem

II, PDS, phytoene desaturase; ZDS, ζ-carotene desaturase. Modified from Carol and

Kuntz (2001).

Regarding its topology, a proteomics approach showed that PDS is

associated to the inner envelope and thylakoid membranes (Joyard et al.,

2009). There is some speculation about the presence of a carotenoid-forming

metabolon allowing substrate channeling. However, supporting experimental

evidence is scarce (Cunningham and Gantt, 1998; Shumskaya et al., 2013;

Nisar et al., 2015). The use of two-dimensional BN-PAGE and western blotting

revealed that PDS is a constituent a 350 kDa complex in chromoplast and

chloroplast membranes (Lopez et al., 2008). However, other possible protein

constituents of the complex were not identified. In the stroma, PDS was part of

a 660 kDa complex, presumably due to chaperon association (Al-Babili et al.,

1996). Lundqvist et al. (2017) applied 1D BN-PAGE followed by mass

Cytb6f

PSIINDH

NADPH NADP+

carotene

product

carotene

substrate

Page 27: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

13

spectrometry. In a chloroplast envelope fraction of Arabidopsis, comigration

of PDS and ZISO in a 75 – 200 kDa fraction and of ZDS and CRTISO in a 200 –

350 kDa fraction was observed suggesting heteromeric complexes. This might

indicate metabolon formation between these enzymes. Similar data obtained

with cyanobacteria support this notion (Takabayashi et al., 2013). Moreover,

gel permeation chromatography, negative staining electron microscopy and

protein X-ray crystallography of soluble OsPDS-His6 in combination with activity

assays suggested the enzyme to be flavinylated and active as homotetramer,

with monomers being deflavinylated and irreversibly inactivated

(Gemmecker et al., 2015; Fig. 1-3 B). However, it remained unclear whether

the homotetramer represents the enzyme’s active state at membrane

surfaces or occurs artifactually.

Investigations carried out in the 1990s revealed differences between bacterial

and plant carotene desaturation. In bacteria, the 4-step carotene desaturase

CrtI (1.3.1.2) introduces all four double bonds yielding lycopene from

phytoene (Linden et al., 1991). PDS as a 2-step desaturase yielded cis-ζ-

carotene from phytoene and was found to be unrelated to CrtI (Pecker et al.,

1992; Linden et al., 1991; Linden et al., 1993). The therefore anticipated plant

ζ-carotene desaturase (ZDS) and its cyanobacterial ortholog CrtQb were later

identified and found to be homologous to PDS (Albrecht et al., 1995 and

Breitenbach et al., 1998, respectively). Compared to PDS, ZDS is much less

studied since sequence homology with PDS and the common introduction of

two double bonds indicated functional analogy. ZDS converts 9,9’-di-cis-ζ-

carotene via identical desaturation reactions at both half sides of the

symmetrical substrate. Double bonds are introduced in cis-configuration at

the positions C7 and C7’, yielding 7,9,9’-tri-cis-neurosporene

(proneurosporene) and 7,9,7’,9’-tetra-cis-lycopene (prolycopene) (Fig.1-5;

Bartley et al., 1999; Breitenbach and Sandmann, 2005). Protein alignments

showed a conserved N-terminal Rossman fold indicating the binding of a

(di)nucleotide cofactor (Breitenbach et al., 1998). Binding of FAD as a redox

cofactor was shown only recently with CrtQb, the ZDS homolog from Nostoc

Page 28: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

14

(Breitenbach et al., 2013). Accordingly, PQ was shown to be the terminal

electron acceptor of ZDS to reoxidize FAD (Breitenbach et al., 1999;

Breitenbach et al., 2013), as already concluded from earlier experiments

(Beyer et al., 1989; Mayer et al., 1990). ZDS activity is consequently thought to

depend on redox regulation via the PQ pool, like PDS (Fig.1-4; Carol and

Kuntz, 2001; McDonald et al., 2011).

Fig.1-5 Scheme of the ZDS reaction with the involved carotene stereoisomers.

9,9’-di-cis-ζ-carotene is converted by ZDS to yield the end product 7,9,7’,9’-tetra-cis-

lycopene via the intermediate 7,9,9’-tri-cis-neurosporene. Desaturation sites are

highlighted in red. The enzyme-bound redox reactions involving FAD and

plastoquinone as well as redox chains that influence the supply of PQ and PQH2 and

thus ZDS activity are depicted. PQ, plastoquinone; PQH2, plastohydroquinone; hν,

light energy (photosynthetic electron transport chain); NDH, NAD(P)H:quinone

oxidoreductase; PTOX, plastid terminal oxidase.

9,9‘-di-cis-ζ-carotene

7,9,9‘-tri-cis-neurosporene

FADoxPQ

PQH2

hνNDH

PTOX

2e-

FADred

FADoxPQ

PQH2

hνNDH

PTOX

2e-

FADred

7,9,7‘,9‘-tetra-cis-lycopene

7

8

8‘7‘

Page 29: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

15

Regarding cis-trans configuration of its substrate, ZDS has special

requirements. Observations in daffodil chromoplasts (Beyer et al., 1989; Mayer

et al., 1990) demonstrated that the 15-cis double bond in the ζ-carotene, as

formed by PDS, needs to be isomerized to trans for ZDS activity (see also

1.3.2). Prolycopene and proneurosporene can only be formed from 9,9’-di-cis-

ζ-carotene in vivo (Bartley et al., 1999) and in vitro (Breitenbach et al., 2005).

9-cis-ζ-carotene was converted to 7,9-di-cis-neurosporene (Breitenbach et al.,

2005) and 9-cis-neurosporene to 7,9-di-cis-lycopene (Bartley et al., 1999). It

was concluded that ZDS recognizes its substrates by the 9-cis configuration of

one half side and can only introduce a 7-cis double bond in the same half

side (Breitenbach et al., 2005). In summary, there are clear similarities

between PDS and ZDS. The main differences are in the regio-specificity of

desaturation and the stereo-configuration of the introduced double bonds.

ZDS introduces cis-double bonds at C7-C8 and C7’-C8’ whereas PDS

introduces trans-double bonds at C11-C12 and C11’-C12’ and simultaneously

isomerizes C9-C10 and C9’-C10’ from trans to cis. It is therefore obvious that

both enzymes must differ mechanistically despite of their overall similarity.

Regarding its localization and topology, a proteomics approach showed that

ZDS is associated with the inner envelope membrane of chloroplasts (Joyard

et al., 2009). Some experiments indicate that ZDS and CRTISO form a

carotenogenic metabolon (Lundqvist et al., 2017). Breitenbach et al. (1999)

suggested dimerization of Capsicum annuum ZDS using gel permeation

chromatography but stated that dimerization was not necessary for ZDS

activity and may be due to unspecific hydrophobic aggregation.

1.3.1.2 Bacterial phytoene desaturase CrtI

Based on sequence comparison, the phytoene desaturase CrtI of bacteria

and fungi appears as evolutionarily unrelated to plant-type PDS. Some

similarities do exist but are largely restricted to the common N-terminal

Rossman fold to bind (di)nucleotide cofactors (Pecker et al., 1992). In

bacteria, there are CrtI enzymes introducing three to six double bonds (3- to

Page 30: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

16

6-step phytoene desaturases; Sandmann, 2009). Only the 4-step CrtI enzymes

yielding lycopene will be discussed here.

CrtI from Pantoea ananatis (formerly Erwinia uredovora) was thoroughly

investigated by Schaub et al. (2012). It catalyzes the sequential desaturation

of 15-cis-phytoene at C11-C12, C11’-C12’, C7-C8 and C7’-C8’, yielding the

end product all-trans-lycopene while only traces of the (presumably) trans-

configured intermediates phytofluene, ζ-carotene, neurosporene are

released (Linden et al., 1991; Fraser et al., 1992). CrtI is thus capable of

performing the entire all-trans carotene desaturation pathway (Fig.1-2). The

enzyme can be isolated as soluble protein but was shown to associate

spontaneously to liposomal membranes representing a monotopic

membrane protein (Schaub et al., 2012). It needs to bind FAD during

membrane association for activity (Schaub et al., 2012; Fraser et al., 1992), a

behavior that is similar to PDS (Al-Babili et al., 1996). FAD serves as the sole

non-exchangeable intermediary electron acceptor which is reoxidized by the

terminal electron acceptor oxygen. This defines CrtI as phytoene:oxygen

oxidoreductase. However, the enzyme was capable of utilizing quinones as

alternative terminal electron acceptors under anaerobic conditions. This

might allow redox regulation via bacterial photosynthetic redox chains

(Schaub et al., 2012). Apart from carotene desaturation, CrtI possesses cis-

trans-isomerase activity under anaerobic conditions, an activity that is

dependent on FADred. The formation of 7,9-di-cis-lycopene from prolycopene

implied half side recognition of carotenoid substrates (Schaub et al., 2012)

which is in accordance with the activity of its homolog, the plant carotene

cis-trans isomerase CRTISO (Isaacson et al., 2002; Isaacson et al., 2004; Yu et

al., 2011).

As CrtI cannot be purified in its flavinylated form (see above), the structure of

the CrtI apoprotein was resolved followed by in silico docking to identify the

position of the isoalloxazine moiety of FAD and of carotene fragments

(Schaub et al., 2012). The CrtI substrate cavity was lined with hydrophobic

residues and accommodated C18 – C20 truncations of phytoene, ζ-carotene

and lycopene which is consistent with a half side recognition of carotenes.

Page 31: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

17

The authors hypothesized that CrtI is active as a dimer that may form on

membrane surfaces possibly upon the binding of a phytoene half side. The

dimer would sequentially introduce all four double bonds after the

isomerization of the central C15-C15´double bond to trans and before

product release (Schaub et al., 2012), as witnessed by the release of only

traces of trans- intermediates. The model of Schaub et al. (2012) is in

accordance with the findings of Raisig and Sandmann (2001). Regio-

specificity of CrtI desaturation was maintained with a C30 15-cis-phytoene that

is truncated by five carbon atoms on both sides, showing that the length of

the substrate half sides is not the sole determinant for regio-specificity. The

central cis-configured triene might be used for the correct positioning of the

substrate (Sandmann, 2009). Site directed mutagenesis revealed that three

conserved histidines at the end of the substrate cavity help determine regio-

specificity of CrtI desaturation (Schaub et al., 2012).

Regarding the reaction mechanism of carotene desaturation, Schaub et al.

(2012) pointed out an aspartate and two arginine residues in active center

acting as a catalytic triad for the acid-base initiated activation of a hydride

that is transferred onto the redox cofactor FAD.

1.3.1.3 Bleaching herbicides targeting PDS

Bleaching herbicides lead to lethal photooxidative destruction of plant cells,

for instance by targeting PDS and ZDS, i.e. carotenogenesis (for review, see

Sandmann et al., 1991; Sandmann and Böger, 1997 and references therein).

Their mode of action, target selectivity and inhibitory effectiveness have been

investigated since the 1980s, however, the native purification of the targeted

carotene desaturases has seldomly been achieved (see 1.3.1.1 and 1.3.1.2).

Consequently, herbicide research relied on observations made in vivo or in

complex cell-free in vitro assays. Considering the interference of, inter alia,

cell structures, cofactors and redox chains with the examined enzymes and

inhibitors, the reliability of such findings may be doubtful. Nowadays, the

emergence of resistances towards widely used herbicides requires a new

focus on herbicide research. This calls for basic research addressing the

Page 32: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

18

native purification of PDS and ZDS and elucidating the inhibitory modes of

action, binding sites, target selectivity and effectiveness of herbicides in order

to possibly design novel agronomically valuable bleaching herbicides.

The most important groups of carotene desaturase inhibitors have been

reviewed by Sandmann (1991). Amongst them, there is a structurally

heterogenous group of PDS inhibitors such as norflurazon (NFZ), fluridone,

diflufenican, fluortamone and fluorochloridone. They share a meta-

trifluoromethylphenyl (m-CF3-phenyl) moiety that is often fused to a second

(heterocyclic) ring structure with a carbonyl group (Babczinski et al., 1995;

Fig.1-6). Among these, the most extensively studied inhibitor is NFZ.

Fig.1-6 Structures of bleaching herbicides targeting PDS.

PDS inhibitors share common structural features as described by Babczinski et al.

(1995), namely a meta-trifluoromethylphenyl moiety that is mostly fused to a second

(heterocyclic) ring structure with a carbonyl group.

NFZ was reported to inhibit PDS non-competitively regarding phytoene

(Sandmann et al., 1989; Mayer et al., 1989) and competitively regarding

plastoquinone, suggesting that NFZ is to be regarded as a structural analog of

PQ (Breitenbach et al., 2001). The mode of action of other m-CF3-phenyl

containing inhibitors has not been examined, except fluridone that was

shown to be non-competitive towards phytoene (Kowalczyk-Schröder et al.,

fluridone

O-(2-phenoxy)ethyl-N-

aralkylcarbamate

fluorochloridone

3-trifluoromethyl-

1,1’-biphenyl

diflufenicannorflurazon

flurtamone trifluoromethyl-phenyl-

ketomorpholine

Page 33: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

19

1992). Flurtamone, fluorochloridone and diflufenican were shown to be more

effective PDS inhibitors than NFZ (Mayer et al., 1989; Sandmann, 1990;

Sandmann and Fraser, 1993; Laber et al., 1999). At the same time, they were

reported to barely impair ZDS activity (Mayer et al., 1989).

Several studies were carried out in the 1980s and 1990s to investigate

structure-activity correlations. NFZ derivatives that had the m-CF3 substituent

replaced by less lipophilic and/or electron-withdrawing substituents showed

substantially decreased inhibition in daffodil chromoplasts and

cyanobacterial membranes (Mayer et al., 1989; Sandmann et al., 1989).

Accordingly, more lipophilic and electron-withdrawing substituents improved

in vivo inhibition in the green algae Scenedesmus obliquus (Sandmann and

Böger, 1982). Quantitative structure-activity correlations were reported with

fluridone derivatives as well (Sandmann et al., 1992) and pointed once more

towards the special importance of the m-CF3 substituent. Additional

important features reside in the second (heterocyclic) ring structure, such as a

carbonyl group in vicinity to the m-CF3-phenyl and the overall lipophilicity and

electronegativity of additional ring substituent (e.g. fluorine atoms).

Cramp et al. (1987) examined different diflufenican derivatives containing a

phenoxy ring replacing the phenyl group in NFZ. They corroborated the results

of Sandmann et al. (1992) regarding lipophilicity and electronegativity of m-

CF3 and demonstrated that the meta position was essential. Sandmann and

Mitchell (2001) examined ketomorpholine herbicides, inhibiting PDS non-

competitively regarding phytoene and consisting of m-CF3-phenyl fused to a

heterocyclic ketomorpholine with a carbonyl (Fig.1-6). The distance between

the carbonyl of the ketomorpholine and m-CF3-phenyl appeared decisive for

inhibitory effectiveness. Ohki et al. (2003) investigated the so-called O-(2-

phenoxy)ethyl-N-aralkylcarbamates, containing a CF3-substituted phenyl

group with a vicinal phenoxy group (Fig.1-6). They emphasized that the

phenoxy group was essential which is in line with the findings on the relevance

of the carbonyl group in other inhibitors. Using daffodil chromoplasts, Laber et

al. (1999) investigated 3-trifluoromethyl-1,1’-biphenyl herbicides. These

contain a m-CF3-phenyl moiety fused to an oxygenated phenyl ring (Fig.1-6).

Page 34: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

20

In accordance with Mayer et al. (1989) and Sandmann et al. (1989), the

replacement of the CF3 group by less lipophilic and electronegative methyl or

hydrogen decreased PDS inhibition. These recurrent principles led to the

assumption that m-CF3-phenyl herbicides all bind to the same site in PDS

(Laber et al., 1999), which was recently shown to be the plastoquinone

binding site for NFZ (Breitenbach et al., 2001; Gemmecker et al., 2015;

Brausemann et al., 2017). This notion finds further support with PDS mutants

showing cross-resistance to these herbicides (see below). Laber et al.

postulated that the binding site consists of a lipophilic pocket

accommodating the m-CF3-phenyl residue, a functionality donating or

accepting hydrogen bonds in vicinity to the m-CF3-phenyl and a lipophilic

area rotated by 30 ° against the plane of the m-CF3-phenyl.

It is noteworthy that the non-homologous bacterial phytoene desaturase CrtI

is not susceptible to these herbicides, probably due to structural differences

and the lack of PQ utilization (see 1.3.1.1 and 1.3.1.2). Consequently, its

expression confers NFZ resistance in plants (Misawa et al., 1994).

Herbicide research also aims at identifying herbicide-resistant enzymes for use

as selective markers. There are several reports on in vivo mutagenesis and NFZ

resistance screening, mainly in cyanobacteria and algae. Five conserved

amino acids in CrtP and PDS were found to mediate resistance against NFZ

upon mutation (Table 1). However, it is elusive whether these residues directly

interact with NFZ or whether allosteric effects are to be considered

(Chamovitz et al., 1993). Some mutants were reported to have partially lost

enzymatic activity, as witnessed by phytoene accumulation and the lowered

carotenoid content in vivo as well as by lowered activity in vitro (Chamovitz

et al., 1993; Martinez-Férez et al., 1994; Table 1). Nevertheless, the

overexpression of such mutated PDS enzymes mediates NFZ resistance in

transgenic plants (Arias et al., 2006; Wagner et al., 2002). Many of these

mutations mediated cross resistance against several m-CF3-phenyl herbicides

to varying degrees (Chamovitz et al., 1993; Martinez-Férez et al., 1994; Arias et

al., 2006), suggesting they occupy the same binding site (Laber et al., 1999).

Page 35: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

21

Table 1 Mutations of conserved amino acids conferring NFZ resistance to CrtP and

PDS enzymes.

Cell-free assays with cell lysates or membranes as enzyme and phytoene sources

were employed to determine PDS activity and NFZ susceptibility. Resistance factors

represent the ratio of the in vitro IC50 for wild type and mutated enzyme. Numbering

of the corresponding mutation site in Oryza sativa PDS refers to the immature protein

including its N-terminal 87 amino acid transit peptide (Acc. A2XDA1.2). Conserved

residues were identified by global protein sequence alignments using the Blosum 62

matrix. n.d., not determined.

publication organism Corresponding

OsPDS mutation

Resistance

factor

Activity

(% wild type)

Suarez et al. (2014) C. reinhardtii Phe162Val n.d. n.d.

Chamovitz et al. (1993) S. sp. PCC 7942 Arg300Pro 23 24

Martinez-Férez et al. (1994) S. sp. PCC 6809 Arg300Cys 25 n.d.

Martinez-Férez et al. (1994) S. sp. PCC 6809 Arg300Ser 220 n.d.

Martinez-Férez et al. (1994) S. sp. PCC 6809 Arg300Pro 149 n.d.

Arias et al. (2006) H. verticillata Arg300His 2 n.d.

Arias et al. (2006) H. verticillata Arg300Ser 18 n.d.

Arias et al. (2006) H. verticillata Arg300Thr 52 n.d.

Arias et al. (2006) H. verticillata Arg300Cys 29 n.d.

Chamovitz et al. (1993) S. sp. PCC 7942 Leu421Pro 8 91

Chamovitz et al. (1993) S. sp. PCC 7942 Val505Gly 15 83

Steinbrenner and

Sandmann (2006) H. pluvialis Leu538Arg 43 n.d.

Huang et al. (2008) C. zofingiensis Leu538Arg 36 n.d.

Liu et al. (2010) C. zofingiensis Leu538Phe 31 129

Liu et al. (2013) C. reinhardtii Leu538Arg 23 80

Liu et al. (2013) C. reinhardtii Leu538Phe 28 129

Chamovitz et al. (1993) S. sp. PCC 7942 Leu538Arg 76 67

Page 36: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

22

1.3.2 ζ-Carotene isomerase ZISO

As outlined in 1.3.1.1, PDS produces 9,15,9’-tri-cis-ζ-carotene whereas the ZDS

substrate is 9,9`-di-cis-ζ-carotene. Observations in daffodil chromoplasts

indicated that the 15-cis double bond is readily photoisomerized to trans, so

that illumination was assumed to be sufficient for the purpose (Beyer et al.,

1989). Accordingly, photoisomerization of ζ-carotene was generally sufficient

in E. coli coexpressing PDS and ZDS – but was rate-limiting for the pathway flux

towards prolycopene (Bartley et al., 1999). Moreover, carotenogenesis occurs

in plant tissues such as roots and inner parts of fruits that are not exposed to

light. The requirement for a ζ-carotene isomerase (ZISO) became obvious.

Li et al. (2007) characterized a maize mutant pale y9 that accumulates

9,15,9’-tri-cis-ζ-carotene in roots and etiolated leaves. The mutated gene,

responsible for the formation of the 9,9’-di-cis isomer, was named ζ-carotene

isomerase (ZISO). Chen et al. (2010) isolated the ZISO ortholog from the A.

thaliana mutant zic1-1 and showed that the Arabidopsis and maize ZISO

orthologs mediated the isomerization of 9,15,9’-tri-cis-ζ-carotene to 9,9’-di-cis-

ζ-carotene in E. coli engineered to accumulate ζ-carotene (Fig. 7). In maize

and Arabidopsis ZISO mutants, the enzyme was found to be essential for

carotenoid biosynthesis both in green and non-green tissues.

Fig. 1-7 Scheme of the ZISO reaction with the involved carotene stereoisomers.

9,15,9’-di-cis-ζ-carotene is converted by ZISO to yield 9,9’-di-cis-ζ-carotene.

9,15,9‘-tri-cis-ζ-carotene

9,9‘-di-cis-ζ-carotene

Page 37: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

23

Bioinformatic analysis revealed that ZISO is conserved amongst

cyanobacteria, algae and plants and that its closest relative is the nitrite and

nitric acid reductase NnrU, a transmembrane protein involved in bacterial

denitrification (Bartnikas et al., 1997). Accordingly, hydropathicity plots

predicted ZISO to possess several transmembrane helices (Chen et al., 2010).

Interestingly, ZISO is a constituent of the Synechococcus elongatus

carotenoid gene cluster that encodes for the enzymes of the poly-cis

desaturation pathway (Chen et al., 2010). In contrast, ZISO orthologs are

absent in the gene cluster coding for the all-trans desaturation pathway in

Rhodobacter sphaeroides, but its homolog NnrU clusters with genes involved

in denitrification. This adds to the notion that NnrU underwent

neofunctionalization in cyanobacteria.

1.4 Apocarotenoids as novel signaling compounds in plants

Apocarotenoids arise from carotenoids upon (non-)enzymatic oxidative

cleavage of a double bond, yielding a ketone or an aldehyde. The plenitude

of cleavable double bonds (see 1.1) and further modification give rise to this

structurally diverse group of metabolites. In plants, they are enzymatically

formed by carotenoid cleavage oxygenases, including carotenoid cleavage

dioxygenases (CCDs) and 9-cis-epoxy carotenoid cleavage dioxygenases

(NCEDs) (for review, see Walter and Strack, 2011 and Hou et al., 2016). They

serve as colorants, odorants and flavors and can attract pollinators or repel

herbivores (for review, see Walter and Strack, 2011 and Hou et al., 2016).

Additionally, the phytohormones ABA and strigolactones are apocarotenoid-

derived signaling compounds (see 1.2.3).

Indirect evidence has pointed towards additional, hitherto unidentified

apocarotenoid derivatives serving as signals in plant developmental

processes and carotenogenesis feedback regulation (for review, see Hou et

al., 2016 and Tian, 2015). For instance, there is ample evidence for positive

and negative feedback regulation originating from lycopene, β-carotene or

xanthophylls (Arango et al., 2014; Bai et al., 2009; Beyer et al., 2002; Bramley et

Page 38: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

24

al., 2002; Cazzonelli and Pogson, 2010; Cuttris et al., 2007; Diretto et al., 2007;

Giuliano et al., 1993; Nogueira et al., 2013; Ronen et al., 2000; Römer et al.,

2000; Römer et al., 2002). However, the sum of observations made with various

plants and tissues are divergent and remain descriptive in the absence of an

identified compound. Apocarotenoids represent prime candidates as they

are less lipophilic and capable of leaving the plastid, especially upon

modification as in ABA and SL biosynthesis.

Two recent publications suggest that apocarotenoid signals originating from

the poly-cis pathway of carotene desaturation regulate carotenogenesis and

plant development (Fig.1-8). If true, this would provide a clue to the

conundrum as to why plants developed the intricate poly-cis desaturation

pathway instead of relying on the simple bacterial trans desaturation

pathway. Avendaño-Vázquez et al. (2014) described a Arabidopsis ZDS

knockout mutant accumulating 9,9’-di-cis-ζ-carotene and traces of 9,15-di-

cis-phytofluene. It showed impaired expression of plastid- and nucleus-

encoded genes and chloroplast biogenesis and possessed needle-like leaves

(Dong et al., 2007). This phenotype was not due to a lack of photoprotection

because it was absent in PSY- or PDS-deficient mutants, devoid of

carotenoids. Blockage of ζ-carotene formation by either PDS inhibition with

fluridone, PDS knockout or PSY knockout alleviated the phenotype of the ZDS

mutant – as well as a AtCCD4 knockout. In contrast, additional knockout

mutants of AtCCD7 and AtCCD8 did not alleviate the ZDS mutant phenotype.

The hypothesis arose that AtCCD4 might cleave putatively cis-configured ζ-

carotene and that the resulting apocarotenoid participates in retrograde

signaling affecting seedling development.

Kachanovsky et al. (2012) examined the carotenoid-deficient yellow flesh

mutant of tomato defective in PSY. Crossing yellow flesh into various CRTISO-

deficient tangerine backgrounds partially restored transcript levels of the

pathway upstream LePSY1 and consequently of carotenoid levels. CRTISO-

deficient tangerine is thus epistatic yellow flesh whereas the ZISO-deficient

zeta is not. The authors concluded that the rescue might be due to the

accumulation of cis-carotenes exclusively accumulating in tangerine but not

Page 39: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

25

in zeta, namely cis-neurosporene or prolycopene. These might be cleaved to

cis-configured apocarotenoids exerting positive feedback regulation on

carotenoid biosynthesis through PSY.

Fig.1-8 The carotenoid biosynthesis pathway in plants and formation of carotenoid

derived signaling molecules.

Desaturation intermediates are shaded in grey. Established and hypothetical signals

(indicated by question marks) that are formed by CCD-mediated cleavage of

desaturation intermediates are given. The formation of abscisic acid and

strigolactones is initiated by NCEDs and CCD7, respectively. GGPP, geranylgeranyl

diphosphate; PSY, phytoene synthase; PDS, phytoene desaturase; ZISO, ζ-carotene

isomerase; ZDS, ζ-carotene desaturase; CRTISO, carotene isomerase; β-LCY,

lycopene-β-cyclase; D27 (DWARF27), all-trans/9-cis-β-carotene isomerase NCED; 9-

cis-epoxycarotenoid cleavage dioxygenase; CCD, carotenoid cleavage

oxygenase. Modified from Bruno et al. (2016).

NCEDs

CCD4 ?

15-cis-phytoene

9,15-di-cis-phytofluene9,15,9‘-tri-cis-ζ-carotene

9,9‘-di-cis-ζ-carotene

7,9,9‘-tri-cis-neurosporene7,9,9‘,7‘-tetra-cis-lycopene

all-trans-lycopene

all-trans-β-carotene

PDS

ZISO

ZDS

CRTISO

β-LYC

abscisic acid

PSY

CCD(s)?

cis-apocarotenoid (?)

2x GGPP

cis-apocarotenoid

chloroplast &leaf developmentfeedback

strigolactones

CCD7

9-cis-β-carotene

CCD8

9-cis-epoxy-xanthophylls

D27

Page 40: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

26

The recent findings by Álvarez et al. (2016) add to this notion. AtPSY transcripts

occur in two alternative splice variants, one with a short 5’UTR that is

abundant, e.g. under light stress for immediate PSY translation. The other one

possesses a long 5’UTR that occurs in two conformations and allows PSY

translation to be finely tuned. The proposed regulation mechanism for the

long 5’UTR resembles a bacterial riboswitch acting as a flux sensor (for review,

see Serganov and Nudler, 2013): A feedback signal from carotenoid

biosynthesis, presumably a short chain apocarotenoid, could directly bind to

the 5’UTR and convert it into a translationally non-permissive configuration.

The decrease in PSY levels would reduce the flux through the pathway (for

review, see Álvarez et al., 2016 and references therein). The authors

emphasize that this post-transcriptional feedback mechanism seems to apply

only to species with one PSY gene such as Arabidopsis. No translation

inhibition was found for a full-length 5’UTR of PSY1 from O. sativa, a species

encoding three PSY genes.

1.5 Objectives of this work

1.5.1 Biochemical characterization of PDS and ZDS

1.5.1.1 Kinetic and structural investigations on PDS and ZDS

The purification to homogeneity of native, highly active PDS and ZDS is

required to enable the detailed investigations on both enzymes (see 1.3.1).

Recently, high quality purification of OsPDS-His6 was achieved, allowing

unprecedented activity in a biphasic in vitro assay and preliminary structure

elucidation by protein crystallography (Gemmecker et al., 2015). In the

course of this work, the protein structures were to be further resolved for PDS

and additionally for ZDS (in collaboration with the team of Prof. Dr. Einsle,

Dept. of Chemistry, University of Freiburg). Implications derived from the PDS

structure would require validation employing kinetic investigations including

mathematical modeling as well as site-directed mutagenesis. Revealing

structure-function relationships regarding the mode of membrane interaction,

Page 41: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

27

substrate recognition and specificity, regio-specificity and reaction

mechanism of carotene desaturation and the identification of rate-limiting

steps stood in the center of the planned research. Comparative kinetic

analyses between OsPDS-His6 and OsZDS-His6 were envisioned as the method

of choice to reveal common and distinct features of the two homologous

carotene desaturases. The comparison with the kinetics and structure of the

bacterial phytoene desaturase CrtI (Schaub et al., 2012) was planned to shed

light on the differences between the two distinct carotene desaturation

systems in bacteria and plants.

1.5.1.2 Action mode of bleaching herbicides targeting PDS

As laid out in 1.3.1.3, structurally different PDS inhibitors share a meta-

trifluoromethylphenyl group and a vicinal oxygen-containing functionality.

The lack of structural information about PDS did not allow the targeted design

of herbicides for the optimization of effectiveness and increased water

solubility leading to increased phloem mobility. With the help of kinetic

inhibitor studies and the interpretation of the OsPDS-His6 crystal structure in a

complex with NFZ, we aimed at elucidating amino acid residues that interact

with the meta-CF3 and carbonyl functionality of NFZ. In cooperation with

researchers of the agrochemical company Syngenta, results might create

knowledge to develop valuable herbicides targeting PDS and herbicide-

resistant PDS mutant enzymes.

1.5.2 Elucidation of the function of ζ-carotene isomerase ZISO

At the beginning of this work, ZISO had not been biochemically characterized

and results from in vivo assays carried out in E. coli had left ample room for

interpretation. For instance, it remained to be elucidated whether ZISO was a

bona fide isomerase for the central C15-C15’ double bond in ζ-carotene or a

modulator of the activity of PDS, capable of isomerizing the C9-C10 double

bond. Additionally, questions related to topology, cofactor requirement,

reaction mechanism and substrate specificity also remained to be addressed.

In view of the experimental challenges with transmembrane proteins like ZISO,

Page 42: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Introduction

28

the minimal goal was to establish an in vitro assay based on ZISO

overexpression in E. coli to examine whether ZISO is in fact an isomerase. If

possible native, homogenous purification was to be established in order to

address remaining questions.

1.5.3 Identification of apocarotenoids originating from the poly-cis pathway

of carotene desaturation

Recently published evidence pleads for the participation of apocarotenoids

derived from poly-cis pathway of carotene desaturation in plastid

development and carotenogenesis feedback regulation (see 1.4). Cleavage

of linear cis-carotenes by a CCD enzyme has been suggested based on

mutant analyses, however, such CCD cleavage activity has not been

investigated so far. It is an objective of this work to assay linear poly-cis

carotene cleavage in vitro with AtCCD1, AtCCD4 and AtCCD7 in search of a

yet unidentified apocarotenoid that might exert signaling functions directly or

upon modification. This could provide an answer to the question why the

much simpler CrtI-based all-trans carotene desaturation pathway in bacteria

and fungi (see 1.3.1) has been abandoned during the evolution of

cyanobacteria and plants.

Page 43: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Materials

29

2 Materials

2.1 Chemicals and materials

carotenoids and apocarotenoids

9-cis-lycopene Buchem, Netherlands

9-cis-β-apo-10’-carotenal Dr. Mark Bruno (University Freiburg)

9-cis-β-carotene Carotenature, Switzerland

all-trans-apo-10-lycopenal BASF, Ludwigshafen

all-trans-lycopene Roth, Karlsruhe

all-trans- / 15-cis-phytofluene (1/1) Carotenature, Switzerland

all-trans-β-carotene Sigma-Aldrich, Taufkirchen

all-trans-β-apo-8-carotenal Sigma-Aldrich, Taufkirchen

15-cis-1´,2´,3´,16´,17´-penta-nor-phytoene Buchem, Netherlands

enzymes

Accuprime GC-rich DNA polymerase ThermoFisher, Dreiech

catalase Sigma, Deisenhofen

deoxyribonuclease I (bovine pancreas) Sigma, Deisenhofen

diaphorase, human Sigma, Deisenhofen

glucose oxidase Sigma, Deisenhofen

lysozyme Roth, Karlsruhe

Phusion High-Fidelity DNA polymerase NEB, Schwalbach

restriction enzymes NEB, Schwalbach

T4 DNA ligase NEB, Schwalbach

Z-competent E. coli Transformation buffer set Zymo Research, Freiburg

kits

Pure YieldTM Plasmid Midiprep System Promega, Mannheim

Pure YieldTM Plasmid Miniprep System Promega, Mannheim

ECLTM Western Blotting Detection Reagents GE Healthcare, Freiburg

GFX PCR DNA / Gel Band Purification Kit GE Healthcare, Freiburg

proteins

alcohol dehydrogenase from S. cerevisiae Sigma, Deisenhofen

apoferritin from equine spleen Sigma, Deisenhofen

bovine serum albumin (BSA) Sigma, Deisenhofen

Page 44: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Materials

30

carboanhydrase from bovine erythrocytes Sigma, Deisenhofen

cytochrome c from equine heart Sigma, Deisenhofen

ovalbumin GE Healthcare, Freiburg

β-amylase from sweet potato Sigma, Deisenhofen

chemicals

β-mercaptoethanol Roth, Karlsruhe

2-(N-morpholino)ethanesulfonic acid (MES) Sigma, Deisenhofen

3-((3-cholamidopropyl)dimethylammonio)-1-propanesulfonate (CHAPS)

Roth, Karlsruhe

agarose Roth, Karlsruhe

ammonium persulfate Roth, Karlsruhe

ampicillin Roth, Karlsruhe

Bacto Agar BD Biosciences, USA

BactoTM Tryptone BD Biosciences, USA

BactoTM Yeast BD Biosciences, USA

blue dextran Sigma, Deisenhofen

bradford staining solution Bio-Rad, München

bromophenol blue Serva, Heidelberg

chloramphenicol Duchefa, Netherlands

Coomassie Brillant Blue G250 Serva, Heidelberg

decylplastoquinone Sigma, Deisenhofen

imidazole Sigma, Deisenhofen

isopropyl-β-D-thiogalactopyranoside (IPTG) Roth, Karlsruhe

kanamycin Sigma, Deisenhofen

L-α-phosphatidylcholine, Type IV-S, > 30 % TLC Sigma, Deisenhofen

Midori green Biozym Scientific, Oldendorf

N,N,N’,N’-tetramethylethylendiamin (TEMED) Roth, Karlsruhe

octyl-β-D-glucopyranoside Biomol, Hamburg

Rotiphorese gel 30 Roth, Karlsruhe

streptomycin Duchefa, Netherlands

tris(2-carboxyethyl)phosphine (TCEP) Thermo Scientific, Dreieich

tris(hydroxymethyl)aminomethane (Tris) Sigma, Deisenhofen

Triton X100 Roth, Karlsruhe

Triton X165 Sigma, Deisenhofen

Page 45: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Materials

31

xylene cyanol Sigma, Deisenhofen

xylenol orange Sigma, Deisenhofen

α-tocopherolacetate Sigma, Deisenhofen

growth media

LB Roth, Karlsruhe

2YT Sigma, Deisenhofen

herbicides

norflurazon Sigma, Deisenhofen

fluridone Sandoz, Switzerland

diflufenican Santa Cruz, Heidelberg

All organic solvents were purchased as pro analysi quality from common suppliers, if used for HPLC analysis as HPLC / LC-MS grade. All other chemicals were purchased as pro analysi quality from common suppliers.

2.2 Primers

P1 5’acaaggaccatagcatatggct3’

P4 5’acggccagtgccaagcttca3’

P2A 5’cctgaagaaatgtgtttaaagcaa3’

P2B 5’cctgaagaaaactgtttaaagcaa3’

P2C 5’catcgaagcgaaatatttctgct3’

P2D 5’catcgaagccctatatttctgc3’

P2E 5’ccaaaaaagatagcaagcccagtttc3’

P2F 5’gggataagctccaacaaagatatg 3’

P3A 5’ttgctttaaacacatttcttcagg3’

P3B 5’ttgctttaaacagttttcttcagg3’

P3C 5’agcagaaatatttcgcttcgatg3’

P3D 5’gcagaaatatagggcttcgatg3’

P3E 5’gaaactgggcttgctatcttttttgg3’

P3F 5’catatctttgttggagcttatccc3’

PDS internal RV 5’catggcaaatatcatggagtgttccttc3‘

PDS into tag 5’ggattataaaatgctatctc3‘

pCDF- Mistic FW 5’ctttaataaggagatatacatgttttgtacattttttgaaaaacatca3‘

Page 46: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Materials

32

linker- Mistic RV

5’ctgctctgaaaatacaggttttcgccgccgctgccaccgccaccttctttttctccttcttcagatactgagatg3‘

linker-ZISO FW 5’gaaaacctgtattttcagagcagcggtggcccgggcagcatggcctccgccgtccgtcc3‘

pCDF-ZISO RV 5‘ccgagctcgaattcggatctcagtgatggtgatggtgatg3‘

pCDF-ZISO FW 5’ctttaataaggagatatacatggcctccgccgtccgtcc3‘

Duet Seq RV 5‘atttcgattatgcggccgtg3‘

2.3 Plasmids

plasmid source description

pRiceOsPDS-His6 Dr. Patrick Schaub (University Freiburg)

IPTG-induced expression of OsPDS with C-terminal His6 tag

pRicemutA I see 3.1.5 IPTG-induced expression of OsPDS mutants with C-terminal His6 tag

pRiceOsZDS-His6 see 3.1.6.1 IPTG-induced expression of OsZDS with C-terminal His6 tag

pCDFDuet-Mistic- OsZISO-His6

see 3.1.6.2 IPTG-induced expression of OsZISO with C-terminal His6 tag and N-terminal Mistic tag

pCDFDuet-OsZISO-His6 see 3.1.6.2 IPTG-induced expression of OsZISO with C-terminal His6 tag

pPhytoen_PS Dr. Patrick Schaub (University Freiburg)

constitutive expression of bacterial CrtE, IPP/DMAPP Isomerase and CrtB to yield phytoene

pz-carotene Prof. Dr. Salim Al-Babili (KAUST, Saudi-Arabia)

constitutive expression of bacterial CrtE, IPP/DMAPP isomerase, CrtB and CrtP to yield cis-ζ-carotenes

pThio-Dan1-AtCCD1 Dr. Mark Bruno (University Freiburg)

arabinose-induced expression of AtCCD1 with N-terminal thioredoxin

pThio-Dan1-AtCCD4 Dr. Mark Bruno (University Freiburg)

arabinose-induced expression of AtCCD4 with N-terminal thioredoxin

pThio-Dan2-AtCCD7 Dr. Adrian Alder (University Freiburg)

arabinose-induced expression of AtCCD7 with N-terminal thioredoxin

pThio- AtCCD8 Dr. Adrian Alder (University Freiburg)

arabinose-induced expression of AtCCD8 with N-terminal thioredoxin

pThio- AtD27 Dr. Mark Bruno (University Freiburg)

arabinose-induced expression of AtD27 with N-terminal thioredoxin

pHLT-LeCrtISO Prof. Dr. Joseph Hirschberg (Hebrew University Jerusalem)

arabinose-induced expression of LeCrtISO with an N-terminal HLT tag

Page 47: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Materials

33

2.4 Consumption materials

thin layer chromatography plates silica 60 F254 Merck, Darmstadt

dialysis tubing Spectra/Por1 30kDa MWCO Spektrum Laboratories, USA

concentrator vivaspin2 hydrosart, 30 kDa MWCO Sartorius, Göttingen

TALON® Metal Affinity Resin Clontech, Saint-Germain-en-

Laye, France

GeneRuler 1 kb DNA ladder Fermentas, St. Leon-Rot

PageRulerTM Prestained Protein Ladder Fermentas, St. Leon-Rot

2.5 Bacterial strains

� DH5α (Thermo Scientific, Dreieich): F- Φ80lacZ∆M15 ∆(lacZYA-argF)

U169 recA1 endA1 hsdR17 (rK–, mK+) phoA supE44 λ– thi-1 gyrA96 relA1

� BL21 (DE3) Gro7 (Clontech, France): F- ompT hsdSB(rB-, mB-) gal dcm (DE3) pGRO7 (CamR)

� Tuner (DE3) (Merck, Darmstadt): F- ompT hsdSB (rB- mB-) gal dcm lacY1 (DE3)

� Rosetta (DE3) (Merck, Darmstadt): F- ompT hsdSB(rB- mB-) gal dcm (DE3) pRARE (CamR)

Page 48: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Methods

34

3 Methods

3.1 Nucleic acid methods

3.1.1 Isolation of nucleic acids

PCR products and other DNA fragments were purified from agarose gels and enzymatic reactions using the Illustra GFXTM PCR DNA / Gel Band Purification Kit (GE Healthcare) according to manufacturer instructions. Plasmids were isolated from E.

coli using the kits PureYieldTM Plasmid Miniprep / Midiprep System (Promega) according to manufacturer instructions.

3.1.2 Separation of nucleic acids

PCR products and plasmids were separated by agarose gelelectrophoresis. Gels with an agarose concentration of 1.0 – 1.2 % (w/v) were prepared with TAE buffer (0.484 % (w/v) Tris, 0.114 % (v/v) glacial acetic acid, 0.037 % (w/v) disodium-EDTA, pH 8.0 adjusted with acetic acid). 0.002 % (v/v) of Midori Green was added for DNA visualization. Samples were mixed with 6x DNA loading buffer (30 % (v/v) glycerol, 0.25 % (w/v) xylencyanol, Orange G and bromophenol blue each). Gels were run in TAE buffer at 7 V cm-1. GeneRulerTM 1 kb DNA ladder (Fermentas) served as a size marker.

3.1.3 Sequencing of nucleic acids

DNA sequencing was carried out by GATC Biotech (Constance, Germany) with the appropriate primers given in 2.2.

3.1.4 Polymerase chain reaction

All reactions were carried out with Phusion ® High Fidelity DNA Polymerase (NEB) or AccuPrime GC-rich DNA polymerase (ThermoFisher) for OsZISO amplification according to manufacturer instructions. Melting temperatures were calculated with the Geneious 8.0.2 software.

3.1.5 Site-directed mutagenesis

Site-directed point mutations were introduced into OsPDS-His6 using overlap extension PCR (Heckman and Pease, 2007). For amplification of the PDS fragment 5’ of the mutated site, primers P1 and P2x were used and primers P4 and P3x for the 3’ fragment. “x” stands for the letter assigned to the mutation (Table 2).

Table 2 Primers used for site-directed mutagenesis of OsPDS and resulting mutations

P1 5’acaaggaccatagcatatggct3’

P4 5’acggccagtgccaagcttca3’

P2A 5’cctgaagaaatgtgtttaaagcaa3’ P3A 5’ttgctttaaacacatttcttcagg3’

P2B 5’cctgaagaaaactgtttaaagcaa3’ P3B 5’ttgctttaaacagttttcttcagg3’

P2C 5’catcgaagcgaaatatttctgct3’ P3C 5’agcagaaatatttcgcttcgatg3’

P2D 5’catcgaagccctatatttctgc3’ P3D 5’gcagaaatatagggcttcgatg3’

P2E 5’ccaaaaaagatagcaagcccagtttc3’ P3E 5’gaaactgggcttgctatcttttttgg3’

Page 49: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Methods

35

P2F 5’gggataagctccaacaaagatatg 3’ P3F 5’catatctttgttggagcttatccc 3’

mutA Arg300Thr mutF Phe162Val

mutB Arg300Ser mutG Met277Leu and Met310Leu

mutC Leu538Phe mutH Thr506Phe and Tyr508Val

mutD Leu538Arg mutI Met277Leu

mutE His159Ala

For mutations mutG / mutH / mutI, a mutated fragment of the PDS cDNA was synthesized and cloned into puC57 by GenScript (New Jersey, USA), excised directly 5’ / 3’ of the fragment with either HindIII and BglII (mut H) or NdeI and SphI (mutG / I) and cloned into the pRice expression backbone previously digested with the required restriction enzymes to reconstitute a point-mutated full length PDS cDNA. Besaid primers were also used to verify all sequences and point mutations in all pRiceOsPDS-His6 plasmids.

3.1.6 Gene cloning

3.1.6.1 OsPDS-His6 and OsZDS-His6

The plasmid pRiceOsPDS-His6 was kindly provided by Dr. Sandra Gemmecker and Dr. Patrick Schaub (University of Freiburg). Briefly, the OsPDS (Gen Bank AF049356) sequence was deprived of the 87 amino acid transit peptide sequence, equipped with a 5´ NdeI site and 3´ His6 coding sequence followed by a HindIII site and synthesized by GenScript®. The vector pCrtI-His6, used previously to express the bacterial carotene desaturase CrtI (Schaub et al., 2012), was digested with NdeI and HindIII to remove the CrtI-His6 cassette and replace it by the OsPDS-His6 coding sequence, resulting in the vector pRicePDS-His6. To obtain a vector pRiceOsZDS-His6, the same procedure as described above was performed with rice ZDS (Gen Bank Acc. AP004273.2), based on a sequence which was codon-optimized, synthesized and sequence-verified by GenScript®.

3.1.6.2 (Mistic-)OsZISO-His6

The coding sequence of OsZISO (Gen Bank Acc. AK066126.1) was deprived of the N-terminal 46 amino acid transit peptide sequence and equipped with a 5’ His6 tag. The sequence was synthesized and cloned into the pUC57 plasmid by GenScript®. The OsZISO-His6 sequence was amplified from pUC57 by PCR using the primers linker-ZISO FW and pCDF-ZISO RV (see 2.2). The sequence for Mistic M110 (Gen Bank Acc. AY874162.1) was amplified from Bacillus subtilis genomic DNA (kindly provided by Dr. Hervé Joel Soufo, University of Freiburg) using the primers pCDF-Mistic FW and linker-Mistic RV (see 2.2), thus depriving it from its stop codon. The coding sequences for Mistic and OsZISO-His6 were cloned into NcoI / BamHI – digested pCDFDuetTM-1 plasmid (Novagen) by Gibson assembly (Gibson et al., 2009). The resulting plasmid pCDFDuet-MisticOsZISO-His6 encoded a chimeric ZISO protein fused to Mistic at its N-terminus via a 20 amino acid flexible linker with a TEV protease cleavage site and rich in glycine, serine and proline. The OsZISO-His6 coding sequence was cloned into pCDFDuetTM-1 stand alone by the same procedure using primers pCDF-ZISO FW and pCDF-ZISO RV (see 2.2), resulting in the plasmid pCDFDuet-OsZISO-His6 encoding a ZISO-His6 protein.

Page 50: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Methods

36

3.1.7 Transformation of E. coli with plasmid DNA

E. coli cells were made competent with the Z-competent E. coli Transformation buffer set (Zymo Research, USA) according to manufacturer instructions and were stored at – 80 °C. 50 µl aliquots of cells were thawed on ice for 5 min, 100 ng of plasmid DNA was added and cells were kept on ice for 30 min. Cells were grown at 37 °C for 1 h after the addition of 1 ml LB medium. They were selected on LB agar plates (1 % (w/v) agar in LB) with the required antibiotic over night at 37 °C (ampicillin, 100 µg ml-1; kanamycin, 50 µg ml-1; chloramphenicol, 20 µg ml-1; streptomycin, 20 µg ml-1).

3.2 Protein methods

3.2.1 Protein expression and purification

3.2.1.1 Expression and IMAC purification of OsPDS-His6

Tuner (DE3) E. coli cells were transformed with pRice-PDS-His6 and grown in 2YT-medium containing ampicillin (100 µg ml-1) under agitation (120 rpm) at 37 °C using baffled Erlenmeyer flasks. The expression of PDS-His6 was induced at OD600nm of 0.5 – 0.7 with 0.5 mM IPTG and took place at 15 °C over night. Cells were harvested by centrifugation, frozen in liquid nitrogen and stored at - 20 °C. IMAC purification was carried out keeping samples on ice. 15 g of cells were resuspended in 20 ml lysis buffer (50 mM Tris-HCl pH 8.0, 100 mM NaCl, 1 mM TCEP) and catalytic amounts of DNase I were added. Cells were disintegrated by using a French Pressure Cell at 1380 bar twice. After 10 min of centrifugation at 20,000 x g, 90 ml of lysis buffer were added to the supernatant and the cell lysate was supplemented with 0.25 % (w/v) CHAPS (0.7 CMC). 4.5 ml of TALON® Metal Affinity Resin (Takara Bio Europe / Clontech) was equilibrated in lysis buffer and added. The mixture was incubated for 30 min at 10 °C under over head agitation (10 rpm). The resin was collected by centrifugation for 5 min at 800 x g. The resin pellet was washed twice – with 40 ml of wash buffer 1 (lysis buffer containing 500 mM NaCl and 2 % glycerol) and 40 ml of wash buffer 2 (wash buffer 1 containing 10 mM imidazole-HCl (pH 8.0)). After reequilibration of the resin with 40 ml of lysis buffer, PDS was eluted from the resin on gravity flow columns with 5 ml of elution buffer (lysis buffer containing 150 mM imidazole-HCl (pH 8.0)). This was followed by dialysis at 10 °C against dialysis buffer (lysis buffer containing 10 % (v/v) glycerol) for imidazole removal. The protein could then be stored at - 80 °C for several months in an active state.

3.2.1.2 Expression and IMAC purification of OsZDS-His6

Rosetta (DE3) E. coli cells were transformed with pRiceZDS-His6 and grown in 2YT-medium containing ampicillin (100 µg ml-1) and chloramphenicol (10 µg ml-1) under agitation (120 rpm) at 37 °C using baffled Erlenmeyer flasks. ZDS-His6 expression was induced at OD600nm 0.5 – 0.7 with 0.2 mM IPTG and protein was expressed at 16 °C over night. Cells were harvested by centrifugation, frozen in liquid nitrogen and stored at - 20 °C. IMAC purification was carried out keeping samples on ice. 15 g of cells were resuspended in 20 ml lysis buffer (20 mM Tris-HCl pH 8.0, 100 mM NaCl, 2 % (v/v) glycerol, 10mM MgCl2, 0.05 % (v/v) Triton X-100 (2.5 CMC), spatula tip of DNase I and lysozyme). Cells were disintegrated by using a French Pressure Cell at 1380 bar twice. After 10 min of centrifugation at 20,000 x g, 90 ml of lysis buffer were added to the supernatant. 4.5 ml of TALON® Metal Affinity Resin (Takara Bio Europe / Clontech) was equilibrated in lysis buffer and added. The mixture was incubated for 30 min at

Page 51: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Methods

37

10 °C under over head agitation (10 rpm). The resin was collected by centrifugation for 5 min at 800 x g. The resin pellet was washed with 40 ml of wash buffer (20 mM Tris-HCl pH 8.0, 500 mM NaCl, 10mM MgCl2,5 % (v/v) glycerol, 15 mM imidazole-HCl (pH 8.0)). After reequilibration of the resin with 40 ml of lysis buffer, ZDS was eluted from the resin on gravity flow columns with 5 ml of elution buffer (20 mM Tris-HCl (pH 8.0), 100 mM NaCl, 10mM MgCl2, 5 % (v/v) glycerol and 100 mM imidazole-HCl (pH 8.0)). This was followed by dialysis at 10 °C against dialysis buffer (lysis buffer containing 10 % glycerol) for imidazole removal. The protein could then be stored at - 80 °C for several weeks in an active state.

3.2.1.3 Expression and crude preparation of Mistic-OsZISO-His6

C43 (DE3) E. coli cells were transformed with pCDFDuet-MisticOsZISO-His6 and grown in 2YT-medium containing streptomycin (50 µg ml-1) under agitation at 37 °C using non-baffled Erlenmeyer flasks. The expression of Mistic-OsZISO-His6 was induced at OD600nm 0.5 – 0.7 with 1 mM IPTG and took place at 20 °C over night. Cells were harvested by centrifugation, frozen in liquid nitrogen and stored at - 20 °C. For crude protein preparation samples were kept on ice. 0.25 g of cells were resuspend in 2 ml of lysis buffer (50 mM Tris-HCl (pH 8.0), 100 mM NaCl, spatula tip of DNase I and lysozyme) and disintegrated by two passages through a French Pressure cell at 1380 bar. The cell lysate was clarified by centrifugation at 20000 x g for 1 min and the supernatant as a crude enzyme source was stored on ice and used for in vitro assays within 2 h.

3.2.1.4 Expression and IMAC purification of LeCRTISO-His6

The plasmid pHLT-CRTISO was kindly provided by Dr. Qiuju Yu (University of Freiburg) in order to express LeCRTISO as a HLT fusion protein with His6 tag and to purify it according to Yu et al. (2011) with minor changes. Briefly, pHLT-LeCRTISO was transformed into BL21 (DE3) E. coli cells harboring the plasmid pGro7 (Takara Bio Inc.), grown in 2YT-medium containing ampicillin (100 µg ml-1) and chloramphenicol (10 µg ml-1) in baffled flasks at 37 °C to an OD600nm of 0.5 – 0.8, induced by adding arabinose (0.2 % (w/v)) and IPTG (0.2 mM) and protein expression took place over night at 16 °C. Cells were harvested and stored at - 20 °C. For IMAC purification, samples were kept on ice or at 4 °C. Cells were disintegrated in 1.5 ml lysis buffer per gram of cell mass (25 mM sodium phosphate buffer (pH 7.5), 2.5 mM MgCl2, 300 mM NaCl, 15 % (v/v) glycerol) by two passages through a French Press Cell at 1380 bar. The crude lysate was clarified by centrifugation at 20,000 × g for 10 min and could either be used as a crude protein preparation directly or for IMAC purification. For further purification, Tween 20 was added to the supernatant to a final concentration of 10 x CMC (0.067 % (v/v)) for membrane solubilization. After 30 min of incubation on ice, TALON® was added and binding was allowed for 45 min at 10 rpm in an overhead shaker. The resin was pelleted by centrifugation at 800 x g for 5 min, washed three times with wash buffer (25 mM sodium phosphate buffer (pH 7.5), 2.5 mM MgCl2, 300 mM NaCl, 15 % (v/v) glycerol, 0.02 % (v/v) Tween 20 and 15 mM imidazole (pH 8.0)). Elution of LeCRTISO-His6 was accomplished on small gravity flow columns by the addition of 100 mM imidazole (pH 8) to the wash buffer. The protein was dialyzed against dialysis buffer (25 mM sodium phosphate buffer (pH 6.2), 2.5 mM MgCl2, 300mM NaCl, 15 % (v/v) glycerol) in a 30 kDa MWCO dialysis tube in order to remove imidazole. The protein was then stored at - 80 °C.

Page 52: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Methods

38

3.2.1.5 Expression and crude preparation for AtCCDs

All four AtCCD enzymes were expressed as thioredoxin fusion proteins using the pThio-vector system. The vectors pThio-Dan1-AtCCD1 and pThio-Dan1-AtCCD4 were kindly provided by Dr. Mark Bruno (University of Freiburg; Bruno et al., 2016), the vectors pThio-Dan2-AtCCD7 and pThio-AtCCD8 were kindly provided by Dr. Adrian Alder (University of Freiburg; Alder et al., 2012). pThio-AtCCD plasmids were transformed into BL21 (DE3) E. coli cells harboring the plasmid pGro7 (Takara Bio Inc.). Cells were grown at 37 °C in 50 ml 2YT growth medium supplemented with ampicillin (100 µg ml-1) and chloramphenicol (10 µg ml-1). Protein expression was induced with 0.2 % (w/v) arabinose at an OD600nm of 0.5. Cells were grown for 4 h at 28 °C, harvested and stored at - 20 °C. For crude protein preparation, cell pellets were resuspended in 1 ml modified LEW buffer (50 mM NaH2PO4, 300 mM NaCl, 1 mg ml-1 lysozyme, 1 mM dithiothreitol (pH 8.0)), passed twice through a French pressure cell at 690 bar and centrifuged at 20,000 x g for 5 min. Protein concentration was quantified using the Quick StartTM Bradford Protein Assay (Bio-Rad Laboratories) and adjusted to 20 µg µl-1.

3.2.1.6 Gel permeation chromatography for OsPDS-His6

For gel permeation chromatography (GPC) analysis of PDS, the GPC system Äkta Explorer 10 consisting of Box-900, pH/C-900, UV-900 und P-900 (Pharma Biotech, Uppsala, Sweden) was used. Samples were separated on a Superose 6 10/300 GL column (GE Healthcare) preequilibrated in GPC buffer (50 mM Tris-HCl (pH 8.0), 100 mM NaCl, 150 mM imidazole-HCl (pH 8.0), 5 mM TCEP) at an isocratic flow rate of 0.8 – 1 ml min-1. Protein elution was monitored via absorption at 280 nm, fluorescence of free and protein-bound FAD at 535 nm upon excitation at 450 nm was monitored with a WatersTM 474 Scanning Fluorescence Detector (Waters GmbH). GPC control and data analysis was carried out using the Unicorn 5.0 software (GE Healthcare). Peak area was quantified by multiplying height of the peak and its width at 50 % height.

3.2.1.7 Gel permeation chromatography for OsZDS-His6

For gel permeation chromatography analysis of ZDS, see 3.2.1.6 with the following changes. Samples were separated in GPC buffer (20 mM Tris-HCl (pH 8.0), 100 mM NaCl, 100 mM imidazole-HCl (pH 8.0), 0.5mM TCEP if not stated otherwise) at an isocratic flow rate of 0.8 – 1ml min-1. For determination of native protein size and potential oligomerization, samples were separated on the columns HiLoad 16/60 (fractionation range 10 – 600 kDa), Superose 6 10/300 (fractionation range 5 – 5000 kDa), Superdex 75 10/300GL (fractionation range 3-75 kDa) (all GE Healthcare) that were calibrated with the following proteins / organic compounds: Blue dextran, carboanhydrase from bovine erythrocytes, alcohol dehydrogenase from S.

cerevisiae, β-amylase from sweet potato, cytochrome c from equine heart, apoferritin from equine spleen, albumin from bovine serum (all Sigma) and ovalbumin (GE Healthcare).

Page 53: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Methods

39

3.2.2 Enzymatic assays

3.2.2.1 OsPDS-His6 in vitro

Under standard conditions, 100 µl of liposomes containing 5 nmol of 15-cis-phytoene (prepared according to 3.2.8) were vortexed after the addition of 0,6 µl of methanolic decylplastoquinone (30 mM). Assay buffer (50 mM MES-KOH (pH 6.0), 100 mM NaCl) was added to a final volume of 700 µl. Assays were started by the addition of 25 µg OsPDS-His6 (ca. 10 µl; see 3.2.1.1). The incubation was carried out at 37 °C in the dark for 15 min. Assays were stopped and carotenoids extracted by mixing them with 300 µl of CHCl3/methanol (2:1, v/v) with α-tocopherol acetate (0.1 mg ml-1) as internal standard. After 5 min of centrifugation at 20000 x g, the carotenoid-containing hypophase was dried, resuspended in 40 µl of CHCl3/methanol (2:1, v/v) and 5 µl were analyzed by HPLC (system 1, see 3.3.1). For inhibitor studies, 2 µl of methanolic inhibitor solution was added to the assay buffer and control assays contained 2 µl of methanol.

3.2.2.2 OsZDS-His6 in vitro

Under standard conditions, 100 µl of liposomes containing 5 nmol of 9,9’-di-cis-ζ-carotene (prepared according to 3.2.8) and were vortexed after the addition of 0.6 µl of methanolic decylplastoquinone (6 mM). Assay buffer (50 mM MES-KOH (pH 6.5), 100 mM NaCl) was added to a final volume of 700 µl. Assays were started by the addition of 25 µg OsZDS-His6 (ca. 10 µl; see 3.2.1.2). The incubation was carried out at 36 °C in the dark for 30 min. Assays were stopped and carotenoids extracted by mixing them with 300 µl of CHCl3/methanol (2:1, v/v) with α-tocopherol acetate (0.1 mg ml-1) as internal standard. After 5 min of centrifugation at 20000 x g, the carotenoid-containing hypophase was dried, resuspended in 40 µl of CHCl3/methanol (2:1, v/v) and 5 µl were analyzed by HPLC (system 2, see 3.3.1). For inhibitor studies, 2 µL of methanolic inhibitor solution was added to the assay buffer and control assays contained 2 µl of methanol. When determining the pH dependency of OsZDS-His6 activity, a combined MES-Tris buffer system (50 mM MES-Tris, 100 mM NaCl, pH adjusted by MES-Tris titration) was used.

3.2.2.3 Mistic-OsZISO-His6 in vitro For in vitro activity measurements, 100 µl of liposomes containing 5 nmol of 9,15,9’-tri-cis-ζ-carotene (prepared according to 3.2.8) were mixed with 100 µl of Mistic-OsZISO-His6–containing cell lysate (see 3.2.1.3) and assay buffer (50mM MES-KOH (pH 6.0), 100 mM NaCl) was added to a final volume of 700 µl. Assays were incubated for 2 h at 30° C in the dark and carotenoids were extracted by mixing samples with 1 volume of CHCl3/methanol (2:1, v/v) with α-tocopherol acetate (0.1 mg ml-1) as internal standard. After 5 min of centrifugation at 20000 x g, the carotenoid-containing hypophase was dried, resuspended in 40 µl of CHCl3/methanol (2:1, v/v) and 5 µl were analyzed by HPLC (system 1, see 3.3.1).

3.2.2.4 LeCRTISO-His6 in vitro

In vitro assays were based on crude enzyme preparation (3.2.1.4). 100 µl of substrate-containing liposomes (prepared according to 3.2.8) were mixed with 191 µl of crude protein preparation and 3 µl of FAD (5 mM). Anaerobic conditions were achieved by adding 3 µl of NADH (5 mM) and freshly prepared sodium dithionite (100 mM) each, resulting in a final volume of 300 µl. Assays were incubated at 37 °C for 1 – 3 h in the dark, depending on the carotenoid substrate. Carotenoid extraction was achieved

Page 54: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Methods

40

by the addition of 900 µl acetone, 300 µl of petroleum ether/diethyl ether (2:1, v/v) with α-tocopherol acetate (0.1 mg ml-1) as internal standard and sonication (Branson Digital Sonifier, USA). After 5 min of centrifugation at 20000 x g, the carotenoid-containing hypophase was dried, resuspended in 40 µl of CHCl3/methanol (2:1, v/v) and 5 µl were analyzed by HPLC (system 6, see 3.3.1). Prolycopene served as a positive activity control.

3.2.2.5 AtCCDs in vitro

Assays were performed in a total volume of 200 µl. Purified substrates were dissolved in CHCl3 (equaling 30 µM as final concentration if not stated otherwise) and mixed with 20 µl ethanolic Triton X-100 (2 % (v/v)) in the case of AtCCD 4 / 7 / 8 and with 50 µl ethanolic ß-octylglucoside (4 % (v/v)) in the case of AtCCD1 dried using a vacuum evaporator and dissolved in 100 µl of 2x incubation buffer (2 mM TCEP, 0.4 mM FeSO4, 200 mM HEPES-NaOH (pH 7.8)and 2 mg ml-1 catalase (Sigma). For 9-cis-lycopene micelles, the substrate was mixed with 10 µl of ethanolic Triton X165 (3 % (v/v)). Assays were started by the addition of each, 50 µl lysate and H2O and then incubated for 1 h, if not stated otherwise, under shaking (200 rpm) at 30 °C in darkness. For extraction, 400 µl acetone were added, followed by short sonication (Branson Digital Sonifier, USA) and the addition of 600 µl petroleum ether/diethyl ether (2:1, v/v) using α-tocopherol acetate (0.1 mg ml-1) as an internal standard. After centrifugation, the epiphase was dried and redissolved in 40 µl CHCl3. 5 µl of the extract were subjected to HPLC (system 7 if not stated otherwise, see 3.3.1). The following substrates served as positive activity controls: β-carotene (AtCCD4), β-apo-8-carotenal (AtCCD1), 9-cis-β-carotene (AtCCD7) and 9-cis-β-apo-10’-carotenal (AtCCD8).

3.2.2.6 OsZISO-His6 in vivo

DH5α E. coli cells were transformed with the plasmid pζ-carotene (see 2.3; kanamycin) to produce tri-cis-ζ-carotene and di-cis-phytofluene by the expression of, inter alia, phytoene desaturase from Synechococcus sp. PCC 6803. For exclusive phytoene production, E. coli was transformed with the plasmid pPhytoene_PS (see 2.3; kanamycin). Cells were cotransformed with pCDFDuet-MisticOsZISO-His6 or pCDFDuet as a control, grown in LB to an OD600nm of 0.5 – 0.7 at 28 °C and ZISO expression was induced by the addition of 0.2 mM IPTG. Cells were further grown in the dark for 18 h at 28 °C, harvested by centrifugation at 8000 x g for 5 min and carotenoids were extracted in the dark by the addition of 3 times 2 ml acetone and sonication (Branson Digital Sonifier, USA). 2 ml of petroleum ether / diethyl ether (2:1, v/v) were added to the combined acetone extract, filled up to 14 ml with water, partitioned and centrifuged at 4000 x g for 5 min. The epiphase was dried and carotenoid extracts were subjected to HPLC to investigate ζ-carotene isomerization (system 3, see 3.3.1). In order to reconstitute the entire poly-cis desaturation pathway and investigate how well light can substitute for ZISO activity in vivo, cells were cotransformed with the plasmids pRiceOsZDS-His6 (ampicillin) and pCDFDuet-MisticOsZISO-His6 (streptomycin) or pCDFDuet-empty as a control. Cells were grown at 32 °C for 8 h after IPTG induction at OD 0.5 – 0.7, either in the dark or under light (800 µmol m-2 sec-1). Carotenoids were extracted and analyzed as given above.

3.2.3 SDS-PAGE

For protein separation, discontinuous sodium dodecyl – polyacrylamide gelelectrophoresis (SDS-PAGE) was applied. Self-casted gels (8.5 x 8.2 x 0.1 cm)

Page 55: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Methods

41

consisted of a 4 % collecting gel and a 8 % / 12 % separation gel composed as follows:

Separation gel 12 % Separation gel 8 % Collecting gel

acrylamide 30 18 ml 12 ml 2,4 ml

separation buffer (1.88 M Tris-HCl; pH 8.8)

9 ml 9 ml x

collecting buffer (1.25 M Tris-HCl; pH 6.8)

x x 1,5 ml

aqua dd 17,4 ml 23,4 ml 10,8 ml

10 % (w/v) SDS 450 µl 450 µl 150 µl

TEMED 22,5 µl 22,5 µl 15 µl

10 % (w/v) APS 210 µl 210 µl 150 µl

Protein samples were mixed with 3 x protein loading buffer (65 mM Tris-HCl (pH 6.75), 20 % (v/v) glycerol, 10 % (v/v) β-mercaptoethanol, 4 % (w/v) SDS, spatula tip of Coomassie Brilliant Blue (CBB) G250), solubilized at 90 °C for 5 min and separated for 1.5 – 2 h at 230 V and 30 mA. Gels were run in SDS-PAGE running buffer (0.25 M Tris, 2 M glycine, 1 % (w/v) SDS). PageRulerTM Prestained-Protein Ladder (Fermentas, St.Leon-Rot) was used as a size marker if not stated otherwise. Gels were stained using Coomassie staining solution (2.5 % (w/v) CBB G250 in ethanol/H2O/acetic acid (227:227:46; v/v/v)) and destaining solution (ethanol/H2O/acetic acid (30:60:10; v/v/v) and scanned.

3.2.4 Protein precipitation

3.2.4.1 Ammonium sulfate precipitation

31 % (w/v) ammonium sulfate was added to PDS or ZDS samples originating from GPC analyses (see 3.2.1.6 and 3.2.1.7), equaling 60 % saturation at 4 °C. Samples were incubated on ice on a rotary shaker (50 rpm) for 30 min and then centrifuged at 20000 x g for 20 min at 4 °C. Protein pellets were resuspended in appropriate volumes of GPC buffer and centrifuged at 20000 x g for 10 min at 4 °C to remove aggregates and denatured protein.

3.2.4.2 Chloroform-methanol precipitation

Proteins samples were denatured and precipitated according to Wessel and Flügge (1984) and resolubilized at 90 °C in 3 x SDS loading buffer for SDS-PAGE (see 3.2.3).

3.2.5 Protein quantification

3.2.5.1 Bradford assay

Protein concentrations were determined using the Quick StartTM Bradford Protein Assay (Bio-Rad Laboratories) according to the manufacturer’s protocol.

Page 56: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Methods

42

3.2.5.2 Nanodrop

IMAC-purified protein was quantified in its dialysis buffer with a Nanodrop photometer (Implen, Munich) using ε280nm = 72,400 l mol-1 cm-1 for OsPDS-His6 and ε280nm = 68,300 l mol-1 cm-1 for OsZDS-His6, both extinction coefficients calculated for the fully reduced proteins with the ExPASy protein parameter online tool.

3.2.6 Membrane association assay

Standard PDS in vitro assays with 50 µg OsPDS-His6 each (see 3.2.2.1) were incubated for 15 min at 37 °C, layered on top of a 30 % (w/v) sucrose in PDS assay buffer (50 mM MES-KOH (pH 6.0), 100 mM NaCl) and centrifuged for 30 min at 110000 x g. The liposomes were isolated from the density boundary, PDS was precipitated by CHCl3/methanol precipitation (see 3.2.4.2) and SDS-PAGE was carried out (see 3.2.3). To test if PDS interaction with liposomes is based on ionic interactions, isolated liposomes were washed with PDS assay buffer supplemented with 0.5 M KCl, incubated for 15 min and PDS liposomes were isolated as described above.

3.2.7 Preparation of carotenoid-containing liposomes

For liposome preparation for in vitro assay with OsPDS-His6, OsZDS-His6 or MIstic-OsZISO-His6 , 5 mg phosphatidylcholine were dissolved in CHCl3 and added to 50 nmol of the required carotenoid substrate. After vortexing, the lipid-carotene mixture was dried under N2 and 1 ml liposome buffer (50 mM Tris-HC (pH 8.0), 200 mM NaCl) was added, followed by 15 min incubation on ice. Liposomes were formed by gentle sonication (Branson Digital Sonifier, USA). Small unilamellar vesicles were formed by a passage through a French pressure cell at 1380 bar. For liposome preparation for CRTISO in vitro assays, 10 mg phosphatidylcholine ml-1 were used as final concentration and liposome buffer consisted of 25 mM sodium phosphate buffer (pH 7.5), 2.5 mM MgCl2, 300mM NaCl and 15 % (v/v) glycerol.

3.2.8 Photometric quantification of protein-bound FAD

OsPDS-His6 or OsZDS-His6 were denatured in dialysis buffer (see 3.2.1.1 and 3.2.1.2) for 10 min at 80 °C in order to release non-covalently bound FAD from the flavoenzymes. Protein was pelleted by centrifugation at 20000 x g for 10 min. FAD in the supernatant was quantified photometrically at 450 nm after establishing a FAD calibration curve. The percentage of flavinylation on a monomeric protein base was calculated from the FAD concentration in the supernatant and the initial protein concentration determined according to 3.2.5.2.

3.3 Chromatography

3.3.1 High performance liquid chromatography

Analytical carotenoid samples were dissolved in 40 µl CHCl3:methanol (2:1, v/v) and 2 – 5 µl were analyzed using a Prominence UFLC XR separation module equipped with a SPD-M20A photodiode array detector (Shimadzu). The column temperature was 40 °C. Peak area integration and data analysis was carried out using the Shimadzu software LabSolutions. For quantification of carotenoids, the HPLC system was calibrated with β-carotene. For PDS assays, molar extinctions coefficients for the all-trans species were used (Britton et al., 1995): 15-cis-phytoene (285 nm) = 68125 l mol-1 cm-1; 9,15-di-cis-phytofluene (350 nm) = 73300 l mol-1 cm-1; 9,15,9’-tri-cis-ζ-

Page 57: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Methods

43

carotene: 400nm = 138000 l mol-1 cm-1. For ZDS assays, values were estimated based on the coefficients published for the trans-carotenoid and the influence of cis-configuration in lycopene isomers according to Hengartner et al. (1992): 9,9’-di-cis-ζ-carotene (400 nm) = 97000 l mol-1 cm-1; 7,9,9’-tri-cis-neurosporene (433 nm) = 108,000 l mol-1 cm-1; 7,9,7’,9’-tetra-cis-lycopene (440 nm) = 100840 l mol-1 cm-1. All other carotenoids were quantified based on molar extinction coefficients published for the trans-configured species (Britton et al., 1995).

HPLC system 1 for PDS assays and ZISO in vitro and in vivo assays: Samples were separated on a YMC carotenoid C30 column (150 mm x 3 mm, 5 µm; YMC) with the solvent system A: methanol / tert-butyl methyl ether (TBME) (1:1, v/v) and B: methanol / TBME / water (5:1:1, v/v/v). The flow rate was 0.7 ml min-1. Separation started at 60 % A and increased linearly to 86 % A within 6.5 min, followed by 2 min of reequilibration.

HPLC system 2 for ZDS assays: Samples were separated on an Accucore C30 column (150 mm x 3.0 mm, 2.6 µm; Thermo Scientific) with the solvent system A: methanol / TBME (1:1, v/v) and B: methanol / TBME / water (5:1:1, v/v/v). The flow rate was 0.7 ml min-1. Separation started at 40 % A and increased linearly to 80 % A within 8 min, followed by 2 min of reequilibration.

HPLC system 3 for ZISO in vivo assays upon co-expression with PDS and ZDS: Samples were separated on a Nucleosil C18 100-5 (150 mm x 4.6 mm, 5 µm; Macherey-Nagel) with an isocratic flow of acetonitrile at 1.4 ml min-1.

HPLC system 4 for purification of ζ-carotene isomers and 9,15-di-cis-phytofluene from

tangerine tomato: Samples were separated on a YMC Carotenoid C30 column (250 mm x 10 mm, 5 µm; YMC) with an isocratic flow of methanol / TBME (4:1, v/v) at 2.0 ml min-1.

HPLC system 5 for purification of prolycopene and proneurosporene from tangerine

tomato: Samples were separated on a Nucleosil C18 100-10 (250 mm x 4.6 mm, 10 µm; Macherey-Nagel) with an isocratic flow of acetonitrile / water (98:2, v/v) at 1.2 ml min-1.

HPLC system 6 for CRTISO in vitro assays and all-trans- and 9-cis-neurosporene

purification: Samples were separated on a YMC carotenoid C30 column (150 mm x 3 mm, 5 µm; YMC) with the solvent system A: methanol / TBME (1:1, v/v) and B: methanol / TBME / water (5:1:1, v/v/v). The flow rate was 0.75 ml min-1. Separation started at 0 % A and increased linearly to 100 % A within 20 min. Final conditions were maintained for 4 min, followed by 2 min of reequilibration.

HPLC system 7 for analysis of CCD assays with phytofluene, ζ-carotene, prolycopene,

proneurosporene: Samples were separated on a YMC carotenoid C30 column (150 mm x 3 mm, 5 µm; YMC) with the solvent system A: methanol / TBME (4:1, v/v) and B: methanol / TBME / water (30:1:10, v/v/v). The flow rate was 0.6 ml min-1. Separation started at 0 % A and increased linearly to 60 % A within 20 min and to 0 % A within 5 min. Conditions were maintained for 9 min, followed by 6 min of reequilibration.

HPLC system 8 for CCD assays with 9-cis-lycopene and all-trans-lycopene: Samples were separated on a YMC carotenoid C30 column (150 mm x 3 mm, 5 µm; YMC) with the solvent system A: methanol / TBME (4:1, v/v) and B: methanol / TBME / water (5:1:1, v/v/v). The flow rate was 0.6 ml min-1. Separation started at 0 % A and

Page 58: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Methods

44

increased linearly to100 % A within 24 min. Conditions were maintained for 4 min, followed by 6 min of reequilibration.

HPLC system 9 for CCD assays with 9-cis-neurosporene: Samples were separated on a YMC carotenoid C30 column (150 mm x 3 mm, 5 µm; YMC) with the solvent system A: methanol / TBME (4:1, v/v) and B: methanol / TBME / water (30:1:10, v/v/v). The flow rate was 0.6 ml min-1. Separation started at 0 % A and increased linearly to100 % A within 20 min. Conditions were maintained for 4 min, followed by 6 min of reequilibration.

3.3.2 Extraction and purification of carotenoids from tangerine tomato and

carotenoid-containing bacteria

Total carotenoid extraction from from bacteria or plants: Either bacterial cell pellets or tomato fruit tissue that was previously homogenized in a blender was extracted in the dark with acetone under sonication (Branson Digital Sonifier, USA) until sufficient decoloration was achieved. The acetone extract was mixed with 1/10 volume of petroleum ether / diethyl ether (2:1, v/v) in a separating funnel. Water was added for partitioning of the carotenoids into the epiphase until a clear phase separation was achieved. The epiphase was collected and dried under vacuum, dissolved in an appropriate volume of CHCl3/methanol (2:1, v/v) and subjected to chromatography required for isolation of the desired carotenoid as follows.

Purification of phytoene from transgenic E. coli containing the plasmid pPhytoene_PS

(see 2.3): The phytoene extract from E. coli was applied on a thin layer silica gel plate 60 F254 (Merck) and developed sufficiently with petroleum ether / diethyl ether (4:1, v/v). The running front containing the phytoene was scrapped and phytoene was eluted with five times 2ml acetone. The eluate was dried under vacuum and phytoene was stored in petroleum ether at - 20 °C in the dark.

Purification of ζ-carotene isomers from transgenic E. coli containing the plasmid pz-

carotene (see 2.3): The extract from ζ-carotene producing E. coli was applied on a silica TLC plate and developed sufficiently with petrolether / diethylether / acetone (4:1:1, v/v/v). The running front, containing the ζ-carotene (and its precursors phytoene and phytofluene), was scrapped and ζ-carotene was eluted from the silica with five times 2 ml acetone. The eluate was dried under vacuum and resuspended in CHCl3/methanol (2:1, v/v). ζ-carotene isomers were isolated with HPLC system 4 (see 3.3.1).

Purification of all-trans-neurosporene from Rhodovulum sulfidophilum (Hagemann et

al. (1996); kindly provided by Nasser Gad’on, University of Freiburg, Microbiology): The carotenoid extract was separated with HPLC system 6 (see 3.3.1) and neurosporene was collected.

Purification of carotenes from tangerine tomato: Total carotene extracts from tangerine tomato (kindly provided by Prof. Dr. Joseph Hirschberg, Hebrew University Jerusalem, Israel) fruits were applied on a silica TLC plate and developed sufficiently in pure petroleum ether. Three fractions were scrapped and eluted with acetone, listed by weaker adsorption: 1) a colorless 9,15-di-cis-phytofluene fraction identified by green fluorescence upon excitation with 350 nm UV light, 2) a bright yellow fraction of 9,9’-di-cis-ζ-carotene below fraction 1 and 3) a dark orange proneurosporene / prolycopene fraction close to the starting point. 9,15-di-cis-phytofluene and 9,9’-di-cis-ζ-carotene were further purified using HPLC system 4 (see

Page 59: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Methods

45

3.3.1). Proneurosporene and prolycopene were further purified using HPLC system 5 (see 3.3.1). The isomeric state of these carotenoids in tangerine tomato fruit was reported by Clough and Pattenden (1979) using nuclear magnetic resonance analysis.

3.3.3 Effective liposomal concentrations of carotenes, decylplastoquinone

and norflurazon

In the PDS assay established by Gemmecker et al. (2015), liposomes are used to accommodate the highly hydrophobic substates phytoene during liposome formation. The second substrate decylplastoquinone and the inhibitor NFZ are added after liposome formation from organic solvent solutions (see 3.2.2.1), both being hydrophobic and partially partitioning into the lipid layer. It is assumed that PDS as a monotopic membrane protein only gains access to its substates and NFZ via the lipid bilayer. Therefore, effective concentrations of besaid compounds – meaning their concentration in the lipid bilayer volume of liposomes (ceff) – are relevant. Therefore, in the framework of the dissertation of Dr. S. Gemmecker (University of Freiburg, 2015) incorporation efficiencies into liposomes were determined according to procedures of Degli Esposti et al. (1983). The incorporation efficiencies were: 100 % for carotenes, 55 % for DPQ and 86 % for NFZ. Given the partial volume of phosphatidylcholine (PC) of 0.997 ml g-1 (Greenwood et al., 2006) and the presence of 0.5 mg PC per assay (see 3.2.2.1), the compounds are present in a lipid bilayer volume of 0.5 µl. Accordingly, ceff is calculated as follows: ceff [mM] = (compound [nmol] / 0.5 µl) * x, with “n” being the amount of carotene, DPQ or NFZ added into the assay and “x” being the incorporation efficiency of the compound given above.

3.3.4 Liquid chromatography – mass spectrometry

3.3.4.1 Identification of apocarotenoids and carotene deuteration analysis

Non-volatile cleavage products produced by CCDs were identified by LC-MS using a Dionex UltiMate 3000 UPLC coupled to a Q-Exactive mass spectrometer (Thermo Fisher Scientific). Sample separation was achieved with a Hypersil Gold C18 UPLC-column (150 x 2.1 mm i.d., 1.9 µm) and the solvent system A, 0.05 % (v/v) formic acid in H2O and B, 0.05 % (v/v) formic acid in acetonitrile. Initial conditions were 70 % B for one minute, followed by a gradient to 100 % B within four minutes. The final conditions were maintained for ten minutes, all at a flow-rate of 0.5 ml min-1. Ionization of apocarotenoids was achieved with atmospheric pressure chemical ionization (APCI) and analyzed in the positive mode. Nitrogen was used as sheath and auxiliary gas, set to 20 and 10 arbitrary units, respectively. The vaporizer temperature was set to 350 °C and the capillary temperature was 320 °C. The spray voltage was set to 5 kV and the normalized collision energy (NCE) to 35 arbitrary units. For data analysis the TraceFinder (3.1) software and authentic apocarotenoid standards were used.

3.3.4.2 Identification of C35 carotenoids

PDS desaturation products originating from 15-cis-nor-phytoene (15-cis-1´,2´,3´,16´,17´-penta-nor-phytoene) were identified by LC-MS using a Dionex UltiMate 3000 UPLC coupled to a Q-Exactive mass spectrometer (Thermo Fisher Scientific). Sample separation was achieved with a YMC carotenoid C30 column (150 mm x 3 mm, 5 µm; YMC) with the solvent system A: methanol / TBME / water (5:1:1, v/v/v) in 0.1 % (v/v) formic acid and B: methanol / TBME (1:1, v/v) in 0.1 % (v/v) formic

Page 60: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Methods

46

acid. Conditions started at 50 % B, increased linearly to 60 % B within 15 min and to 100 % B within further 5 min. Final conditions were maintained for 10min, all at a flow-rate of 0.6 ml min-1. Ionization of apocarotenoids was achieved with atmospheric pressure chemical ionization (APCI) and analyzed in the positive mode. Nitrogen was used as sheath and auxiliary gas, set to 20 and 10 arbitrary units, respectively. The vaporizer temperature was set to 350 °C and the capillary temperature was 320 °C. The spray voltage was set to 5 kV and the normalized collision energy (NCE) to 35 arbitrary units. For data analysis the TraceFinder (3.1) software and authentic apocarotenoid standards were used.

3.3.4.3 Identification of protein-bound nucleotide cofactors

Sample separation was achieved with a Hypersil Gold C18 UPLC column (150 x 2.1 mm i.d., 1.9 µm) and the solvent system A: 50 mM ammonium acetate in 1 % (v/v) formic acid in H2O and B: 1.7 mM ammonium acetate in H2O. The flow rate was maintained at 0.5 ml min-1. Conditions started at 100 % A, decreased to 50 % A within 10 min and finale conditions were maintained for 5 min. Ionization of mono-/dinucleotide cofactors was achieved with electrospray ionization (ESI) and analyzed in the positive mode. Nitrogen was used as sheath and auxiliary gas, set to 40 and 10 arbitrary units, respectively. The vaporizer temperature was set to 300 °C and the capillary temperature was 350 °C. The spray voltage was set to 3.5 kV and the normalized collision energy (NCE) to 25 / 40 arbitrary units. For analysis of the full MS and data dependant MS2 data the TraceFinder 3.1 software was used. For cofactor identification, the following combinations of precursor ions [M+H]+ in MS1 and fragment ions in MS2 (in brackets) were established with authentic, purified standards: FAD, MS1 m/z 786.16441 (348.1, 439); FMN, MS1 m/z 457.11689 (359.2, 439.1); NAD+, MS1 m/z 664.11640 (524, 542.1); NADH, MS1 m/z 666.13205 (348.2, 649.2); NADPH, MS1 m/z 745.09055 (428.1, 729.1).

3.3.5 Gas chromatography – mass spectrometry

Volatile cleavage products such as 6-methyl-5-hepten-2-one and geranylacetone were collected by solid phase micro extraction (SPME; PDMS, 100 µm; Supelco). The fiber was exposed to the in vitro assay head space for 15 min and thermodesorbed in the injector of the Trace GC coupled to a Trace DSQ II mass spectrometer (Thermo Fisher Scientific). Separation was achieved on a 30 m Zebron ZB-5 column 0.25 mm i.d., 0.25 µm film thickness (Phenomenex). The initial temperature of 50 °C was maintained constant for five minutes, followed by a ramp of 25 °C min-1 to a final temperature of 280 °C which was maintained for five minutes. The helium carrier gas flow rate was 1 ml min-1 and the injector temperature was set to 220 °C. Electron impact ionization (EI) was used at an ion source potential of 70 eV and a source temperature of 200 °C. Spectra were matched to the NIST (2.0) database using the Excalibur software. Additionally, standards of geranylacetone and 6-methyl-5-hepten-2-one were used (Sigma).

3.4 Cryo scanning electron microscopy

The liposomes with bound OsPDS-His6 were isolated from membrane binding assays (see 3.2.6) and concentrated by ultracentrifugation at 150000 x g for 30 min. The enzymatic activity of PDS was verified by the observed yellow color of the liposomes upon conversion of phytoene to ζ-carotene. The isolated liposomes of four PDS in

vitro assays (see 3.2.2.1) with 50 µg PDS each were combined to 40 µl of PDS-

Page 61: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Methods

47

liposome suspension (in 12.5 mM MES-KOH, pH 6.0). After the addition of 30 % (v/v) glycerol, suspension was pipetted into the 50 µm cavity of two 3 mm aluminum specimen carriers before sandwiching them. The assembly was frozen using the HPM 100 (Leica) freezer, transferred into the Freeze Fracture System EM BAF060 (Leica) and fractured. Samples were visualized directly in a Zeiss Auriga SEM system (- 115 °C, 5 kV acceleration voltage, 20 µm aperture using the inlens SE detector) or after 5 min of sublimation at - 105 °C in order to display liposomal surfaces. Sublimated as well as untreated samples were coated with 2.5 nm of platinum/carbon and backed with 4 nm of carbon at a gun angle of 45° and under steady stage rotation with 40 rpm.

3.5 Homology modeling and in silico docking of OsZDS-His6

Homology modeling of Oryza sativa ZDS (XP_015646524.1), deprived of its N-terminal plastid target peptide and equipped with an N-terminal methionine, was carried out using the SWISS-MODEL protein structure homology modeling server in automated mode with the PDS structure as template (Brausemann et al., 2017). The obtained ZDS structure represented the apoenzyme. Riboflavin as a precursor of the ZDS redox cofactor FAD was docked into the ZDS structure using the SwissDock ligand docking web service. The thermodynamically most favorable holoenzyme structure, i.e. the structure with the lowest free enthalpy ΔG, obtained from in silico docking was chosen for structural analysis of OsZDS.

3.6 Mathematical modeling of PDS time courses and kinetics

General procedures. The model consists of a set of ordinary differential equations (ODEs) for the contributing processes following mass action kinetics. The maximum likelihood method is used to estimate parameters such that the model prediction optimally describes the observed PDS time courses. Setting up the likelihood,

normally distributed noise is assumed. The cost function χ�(θ) = ∑ (��(�, ))���� needs to

be minimized in order to maximize the likelihood. Here, θ denotes the model parameters, the index i runs over the data points taken at time ti with value xi and uncertainty σ� and x(t�, θ) is the model prediction at time t� . The nonlinear minimization of the cost function is performed by a trust region optimizer (Nocedal and Wright, 1999). Derivatives of the cost function, upon which the optimizer relies, are provided by sensitivity equations. Prior knowledge about parameter values, e.g. values of the initial states, are incorporated by either fixing the parameter value or adding a penalty to the cost function via a quadratic prior function. In general, the cost function can have several local optima, besides the global optimum. In order to find the global optimum a multistart approach is performed by seeding the optimization in different points of the parameter space. The ODEs and sensitivity equations are integrated with the lsodes solver (Soetaert et al., 2010). Identifiability of the parameters and their confidence intervals are determined by the profile likelihood method (Raue et al., 2009). The model was implemented using the dMod package for dynamic modeling in R (Kaschek et al., 2016). Data preprocessing. For PDS time courses of the conversion of phytoene and phytofluene, the amounts of phytoene, phytofluene and ζ-carotene were measured over time. The experiments were conducted in triplicate. Uncertainties for the computed mean values were first estimated by a maximum likelihood method

Page 62: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Methods

48

combining the empirical mean values and variances with an error model. However, additional fluctuations between neighboring time points, larger than those represented by the replicates, were observed. They cannot be captured by the error model described above, but would lead to an underestimation of the derived parameter profiles and uncertainties. Therefore, the uncertainty parameters of the error model were estimated together with the other model parameters, including the log(σ�)-term originally contained in the log-likelihood, leading to a new cost function:

−2 log L (θ) = ��x�(θ) −x��σ�(θ) ��

�+ log (σ�(θ)�)

The uncertainty parameters σ� include a relative and an absolute contribution for each observable, e.g. σ[!] =σ[!]#$% ∙ [p] + σ[!]()* and may vary between the different

reaction time courses. The relative normalizations of phytoene, phytofluene and ζ-carotene measurements were investigated by a preceding optimization. It is based on conservation of mass, i.e. the total sum of carotenes is conserved during reaction time courses. Such normalization is needed because of inaccuracies during carotene quantification. The molar extinction coefficient is known for 15-cis-phytoene but not for 9,15-di-cis-phytofluene and 9,15,9’-tri-cis-ζ-carotene. Therefore, the molar extinction coefficients for the all-trans species of phytofluene and ζ-carotene are used as approximation. Scaling parameters s,!, s,!- and s,. for phytoene, phytofluene and ζ-carotene,

respectively, were estimated by minimizing the discrepancy s,! ∙ [p]�/� +s,!- ∙ [pf]�/� +s,. ∙ [z]�/� − c

at all time points t� for an arbitrary constant c. Since the absolute scale incorporated

by the constant c is unknown, the ratios l3 = *,4*,5 and ratios l� = *,�

*,5 including their

confidence intervals are estimated by a least squares approach. The scaling parameters s! , s!- and s. used for phytoene, phytofluene and ζ-carotene in the

model prediction are related to the ratios via s! = l� ∙ s!- and s. = %4%� s!- and the

constraints on l3 and l� are added via a quadratic prior to the cost function. For additional information about data preprocessing, see supplemental data.

3.7 Deduction of non-covalent interactions from changes in electron

density gradients

A single point ab initio quantum mechanical calculation (density functional theory,

6-31G* model chemistry) was performed on the complex of interest, the PDS – NFZ

cocrystal structure (Brausemann et al., 2017). Non-covalent interactions were

deduced from changes in the gradient of the electron density between atoms

according to the methods described by Johnson et al. (2010). The system for

calculations consisted of only 16 flavin and PDS residues with a total of 313 atoms

that are in direct contact with norflurazon (irrelevant parts of each residue

disconsidered), representing a pared down representation of the PDS – NFZ binding

site.

Page 63: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

49

4 Results

4.1 Biochemical characterization of phytoene desaturase PDS

4.1.1 Association and oligomerization of OsPDS-His6 at liposomal surfaces

PDS associates with plastid membranes to access its hydrophobic,

membrane-soluble carotene substrates and plastoquinone. As suggested by

hydropathy plots, the protein has no transmembrane helices. Moreover, it can

be natively purified as soluble in the absence of detergents and associates

spontaneously to liposomes to convert phytoene to ζ-carotene (Gemmecker

et al., 2015). Accordingly, PDS is assumed to represent a monotopic

membrane protein, interacting with only one lipid leaflet. Furthermore,

experiments pointed towards homotetrameric assembly of OsPDS-His6 in

solution (Fig. 1-3 B). It remained unclear whether homotetramers represent the

catalytically active form at membrane surfaces.

4.1.1.1 Monotopic association of OsPDS-His6 with liposomes

In order to better characterize the mode of membrane association, the

interaction of OsPDS-His6 with phosphatidylcholine liposomes was examined.

First, it was tested whether membrane interaction required only the

phospholipid bilayer or whether the simultaneous presence of substrates (DPQ

and phytoene) was required (Fig. 4-1 A). As depicted in lane “1 L”, OsPDS-His6

readily attached to phosphatidylcholine liposomes containing the substrates.

A small portion of the protein formed aggregates that appeared in the pellet

upon centrifugation (lane 1 P). In the absence of liposomes, OsPDS-His6

formed aggregates (lane 2 P) since the assay buffer contained neither

imidazole nor glycerol, both promoting protein solubility in the absence of

membranes (Gemmecker et al., 2015). Lane “3 L” shows that OsPDS-His6

readily attached to membranes in the absence of substrates. Second, it was

tested whether PDS membrane association mainly relies on electrostatic

interactions or rather on hydrophobic interactions. Washing proteoliposomes

with a high-salt buffer did not detach PDS, as the amount of protein

recovered was very similar as upon washing with a low-salt buffer (Fig. 4-1 B,

Page 64: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

50

compare lanes 1 and 2). This points towards hydrophobic interaction of the

protein with the membrane and accordingly, PDS could only be released

from liposomes by solubilization with detergents (not shown). The OsPDS-His6

interaction with membranes requires only phospholipid bilayers and is mainly

mediated by hydrophobic interactions with the membrane core.

Fig. 4-1 SDS-PAGE analysis of OsPDS-His6 liposome binding assays.

(A) Dependency of OsPDS-His6 membrane association on substrate-containing

liposomes (lane 1) and substrate-free phosphatidylcholine liposomes (lane 3).

Fraction L represents liposome-bound OsPDS-His6 and fraction P represents pelletable

aggregates of unbound enzyme. Conducting the assay in the absence of liposomes

led to complete OsPDS-His6 precipitation (lane 2). (B) Characterization of the mode

of interaction between OsPDS-His6 and liposomes. Lane 1 represents protein bound

to liposomes washed with 100 mM KCl, lane 2 washed with 500 mM KCl.

4.1.1.2 Homooligomerization of OsPDS-His6 at liposomal membrane surfaces

Cryo scanning electron microscopy (cryo-SEM) was used to address the

question whether the PDS homotetramers found in solution (Gemmecker et

al., 2015; Fig. 1-3 B) are artifactual or rather represent the biologically active

form at membrane surfaces. The size of OsPDS-His6 in its active state at

membrane surfaces should be compared to the size of soluble OsPDS-His6

homotetramers previously investigated by negative staining. Additionally,

liposome fracture faces should reveal the presumed absence of membrane

spans.

21 3

L P L P L P

1 2A BL L

Page 65: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

51

Supporting monotopic membrane interaction, fracture faces of PDS

proteoliposomes capable of performing phytoene desaturation (Fig. 4-2 A, B,

black arrow in Fig. 4-2 B) did not support the presence of membrane spans.

Sublimation revealed the liposomal surface with OsPDS-His6 particles (Fig. 4-2

B, red arrows) that were 14.5 ± 1.9 nm in diameter (n = 30), whereof 2 nm are

due to Pt/C-coating. Accordingly, active OsPDS-His6 at liposomal surfaces is

12.5 nm ± 1.9 nm in size. This corresponds well with the reported size of soluble

homotetramers of 11.8 ± 1.3 nm (Gemmecker et al., 2015).

Fig. 4-2 Freeze fracture electron microscopy of OsPDS-His6 proteoliposomes.

(A) Membrane fracture face of OsPDS-His6 proteoliposome visualized by cryo

scanning electron microscopy. The outer fracture face boundary, surrounded by ice,

is indicated by a black arrow. (B) Membrane fracture face (surface/fracture face

boundary indicated by black arrow) and membrane surface with associated OsPDS-

His6 particles (red arrows) exposed upon ice sublimation.

In summary, cryo-SEM analysis corroborated the notion that PDS is a

monotopic enzyme being present as homotetramer in its active state on

membrane surfaces. In accordance with this, PDS monomers were shown to

be unflavinylated and enzymatically inactive (Gemmecker et al., 2015).

4.1.2 Regio-specificity of carotene desaturation by OsPDS-His6

PDS catalyzes a highly regio-specific reaction. No carotene desaturation at

positions other than C11-C12 and C11’-C12’ has ever been reported. It

remains elusive how the correct positioning of the reaction sites relative to the

PDSA B

200 nm

Page 66: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

52

active center is achieved. In VP14, a 9-cis-epoxy-carotenoid cleavage

dioxygenase, the 9-cis-configured double bond acts as a restrictor arresting

the carotenoid in the correct position relative to the active site (Messing et al.,

2010). Alternatively, the long carotenoid substrate might be introduced into

the substrate cavity until reaching the cavity end that acts as restrictor. In this

case, the substrate length might mediate correct positioning.

The modeling of 15-cis-phytoene in its extended conformation into the

substrate cavity of OsPDS-His6 supports both scenarios (Fig. 4-3). Charged

amino acid residues at the back end of the hydrophobic substrate cavity

might limit phytoene introduction by repulsive forces (Fig. 4-3 B). Notably,

displacement of coordinated water molecules from the cavity would be

required upon phytoene binding to properly align the C11-C12 bond with the

isoalloxazine (Fig. 4-3 B). Alternatively, the bent topology of the substrate

channel in vicinity of the isoalloxazine might arrest the kinked 15-cis double of

phytoene and mediate adequate positioning (Fig. 4-3 A, B).

Fig. 4-3 Hydrophobic substrate cavity of OsPDS-His6.

(A) The hydrophobic channel of OsPDS-His6 is depicted and colored-coded

according to its electrostatic surface potential from - 5 to 5 kbT/e. The channel-

forming residues are shown as sticks, FAD is given as stick and balls. (B) The channel

topography was determined with the Caver 3.0 software. 15-cis-phytoene is shown in

C

BA

Page 67: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

53

its approximate size. The reaction sites of desaturation (C11-C12; C11’-C12’) are

highlighted in red, the sites of isomerization (C9-C10; C9’-C10’) with triangles.

Hydrophobic and hydrophilic cavity sections as well as the position of the redox-

active isoalloxazine of FAD are marked. Water molecules are given as blue circles.

(C) Waters molecules in the charged channel part are shown as spheres and

numbered as in (B). The coordinating amino acids, inter alia Thr508 and Tyr506, are

shown as sticks. Figures are from Brausemann et al. (2017).

4.1.2.1 Regio-specificity with 15-cis-penta-nor-phytoene (C35)

To elucidate the mode of substrate positioning, a C5-truncated variant of 15-

cis-phytoene (15-cis-1’,2’,3’,16’,17’-penta-nor-phytoene; hereafter 15-cis-nor-

phytoene; Fig. 4-4 A) was incorporated into liposomes and used as a

substrate. Assuming the 15-cis-configuration to be decisive (scenario I),

OsPDS-His6 would maintain specificity for C11-C12 and C11’-C12’, irrespective

whether the full length or the truncated half side of 15-cis-nor-phytoene is

introduced into the cavity (Fig. 4-4 A, scenario I). Accordingly, this asymmetric

substrate would be desaturated twice, yielding an end product with a

chromophore and UV/VIS spectrum identical to that of 9,15,9’-tri-cis-ζ-

carotene. Assuming that the cavity back end is crucial (scenario II), the regio-

specificity of desaturation is expected to be disturbed when the truncated

half side of 15-cis-nor-phytoene is introduced (Fig. 4-4 A, scenario IIb). Being

shorter, the C11’-C12’ bond would slip beyond the redox-active isoalloxazine

and instead, the central triene around C15=C15’ would occupy this position.

Consequently, carotene desaturation would be impossible. In contrast, regio-

specificity for C11-C12 would be maintained with the untruncated half and

carotene desaturation could occur (Fig. 4-4 A, scenario IIa). In scenario II, 15-

cis-nor-phytoene could only be desaturated once to yield a product with a

pentaene chromophore identical to that of 9,15-di-cis-phytofluene.

Page 68: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

54

Fig. 4-4 LC-MS analysis of desaturation products formed from 15-cis-nor-phytoene.

(A) Structure of 15-cis-1´,2´,3´,16´,17´-penta-nor-phytoene (15-cis-nor-phytoene, C35).

The desaturation sites C11-C12 and C11’-C12’ and the central C15-C15’ double

bond are labeled. The single bonds to be desaturated, i.e. to be positioned correctly

above the isoalloxazine are indicated by arrows. If substrate positioning is mediated

by the 15-cis double bond (I), two double bonds can be formed. With the substrate

cavity end being decisive, substrate length would be crucial allowing only one

desaturation reaction to take place (II). (B) Product formation from 15-cis-nor-

phytoene was assessed by LC-MS and detected photometrically at 275 – 400 nm

(top panel). The UV/VIS spectra of substrate and products are shown in the central

panel. The bottom panel shows the corresponding MS1 spectra, [M+H]+ masses, the

derived sum formulas and the mass deviation.

285 nm 348 nm366

331

398 nm

423378

295

400 450 500 550m/z

102030405060708090

100

Rel

ativ

e A

bun

dan

ce 477.44563C35H57

0.31184 ppm

0

m/z420 460 500 540

475.43002

0.39311 ppmC35H55

468 472 476m/z

473.41434

0.33579 ppmC35H53

18.21 min 20.68 min 24.06 min

min17 18 19 20 21 22 23 24 25

18.17I

20.63

II24.01m

AU

(275

-40

0n

m)

I*22.36

II*22.92

A

B

1 4‘

11‘12‘12

11

15 15´

I

IIa IIb

15-cis-nor-phytoene

Page 69: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

55

The products formed by OsPDS-His6 from 15-cis-nor-phytoene were analyzed

by LC-MS. The substrate 15-cis-nor-phytoene revealed an UV/VIS spectrum

resembling the one of 15-cis-phytoene and the quasimolecular ion [M+H]+

corresponded to the calculated exact mass of C35H57 (Fig. 4-4 B left). The

molecular mass of the major desaturation product I corresponded to C35H55. It

had thus undergone one desaturation reaction, and its UV/VIS spectrum

strongly resembled the one of 9,15-di-cis-phytofluene, revealing the presence

of a pentaene chromophore (Fig. 4-4 B, middle). The minor product I* with a

similar UV/VIS spectrum and the same molecular mass represents a geometric

isomer of product I. Two additional desaturation products, II and II*, were

formed at lower levels and shared a UV/VIS spectrum strongly resembling the

one of 9,15,9’-tri-cis-ζ-carotene. This indicates the presence of a heptaene

chromophore. The molecular mass of II and II* corresponded to C35H53,

implying two desaturation reactions of 15-cis-nor-phytoene (Fig. 4-4 B right).

Conclusively, the identification of a desaturation product with two introduced

double bonds, i.e. the loss of four mass units for [M+H]+ (from C35H57 to C35H53),

and the ζ-carotene-like heptaene chromophore and UV/VIS spectrum

support that the 15-cis-configuration of phytoene is decisive to attain regio-

specificity in PDS. However, it is to be noted that this is not the main product.

The intermediate with one introduced double bond (C35H55 for [M+H]+)

exhibiting a phytofluene-like UV/VIS spectrum predominates. This may be due

to the truncation hindering substrate binding at this end.

4.1.2.2 Regio-specificity upon mutation of the substrate cavity back end

As depicted in Fig. 4-3 B, Thr508 and Tyr506 reside at the postulated hydrophilic

cavity back end coordinating water molecules. In order to scrutinize if these

residues contribute to the regio-specificity of desaturation (scenario II, see

4.1.2.1), the mutations Tyr506Phe and Thr508Val were simultaneously introduced.

We hypothesized that introducing hydrophobic residues would elongate the

cavity, prevent water coordination and result in erroneous regio-specificity if

the cavity back end was decisive. However, HPLC analysis revealed that 15-

Page 70: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

56

cis-phytoene was converted into canonical 9,15-di-cis-phytofluene and

9,15,9’-tri-cis-ζ-carotene, albeit at lowered activity (< 5% activity of the wild

type enzyme; data not shown). Thus, regio-specificity persisted. This

corroborates the relevance of the cis-configured central double and the

bent substrate cavity topology. The fact that Tyr506 is conserved in PDS

enzymes whereas position 508 can be occupied by hydrophilic as well as

hydrophobic amino acids (see protein alignment in Fig. 4-16) adds to the

notion that the cavity back end is of minor relevance in this respect.

4.1.3 Reaction mechanism of carotene desaturation in OsPDS-His6

The structure of the bacterial phytoene desaturase CrtI, a flavoprotein that is

inhomologous to PDS, has been elucidated and a reaction mechanism was

postulated (Schaub et al., 2012). In short, the involvement a catalytic triad of

acidic and basic amino acid residues was proposed that activated hydride

abstraction and carbocation formation. During the proposed carotene

desaturation mechanism, protons are supplied via the acidic residue, as

corroborated by the incorporation of deuterium from heavy water (D2O).

The analysis of the PDS structure revealed that the FAD-containing active

center lacks functional residues that could form such a catalytic triad. There

are only two charged residues in less than 8 Å distance from the isoalloxazine

that could initiate an acid-base type mechanism of catalysis: His159,

potentially acting both as a base and an acid due to its neutral pK, and the

basic Arg300 (Fig. 4-5 A). Both residues are conserved in PDS enzymes (Fig. 4-

16). However, a direct interaction of His159 and the carotene appears unlikely

since it is not in a direct contact with the substrate cavity but localized

beneath the isoalloxazine (Fig. 4-5 A). The position of Arg300 might allow direct

interaction with the reaction site whereas Lys142, Glu451 and Glu542 are most

likely too distant (Fig. 4-5 A). Arg300 and His159 were exchanged individually for

uncharged amino acids by site directed mutagenesis (Arg300Ser, Arg300Thr,

His159Phe). The mutant enzymes exhibited catalytic activity, albeit at lower

Page 71: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

57

levels. This indicates that Arg300 and His159 are not catalytically essential, i.e.

not involved in acid-base catalysis of carotene desaturation.

This lack of functional active center residues is unprecedented in

flavoproteins, particularly in those catalyzing the formation or rupture of more

than one stable chemical bond. FAD occurs in a plenitude of redox states

and is thus able to mediate a variety of reaction mechanisms. Taking into

account that PDS catalyzes simultaneous trans-desaturation of phytoene at

position C11-C12 and trans-to-cis isomerization at position C9-C10, two “flavin

only” reaction mechanisms for PDS are conceivable. The redox-reactive N1-

N5 isoalloxazine functionalities might act as the sole catalyst in carotene

desaturation (Fig. 4-5 A and B; for details, see Brausemann et al., 2017). Since

acidic amino acids are catalytically involved, no proton incorporation from

external sources into the carotene product is postulated (Fig. 4-5 B).

In order to verify the absence of acid-base mechanisms, phytoene

desaturation was carried out in reaction buffers prepared in D2O and

deuterium incorporation into ζ-carotene was monitored by LC-MS (Fig. 4-5 B,

C). Canonical 9,15,9’-tri-cis-ζ-carotene was formed both in buffers containing

D2O or H2O, as witnessed by the UV/VIS spectra (Fig. 4-6 A), the same [M+H]+

mass of 541.47678 ± 5 ppm corresponding to C40H61 (Fig. 4-6 B, C) and the

same mass for the M+1 signal that is due to 13C incorporation. Thus, unlike the

situation found with CrtI (Schaub et al., 2012), there is no deuterium

incorporation so that an acid-base initiated catalysis of carotene

desaturation by PDS appears improbable. This result and the lack of charged

residues in proximity to the active site rather support a “flavin only” reaction

mechanism.

Page 72: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

58

Fig. 4-5 Active center of OsPDS-His6 and proposed “flavin only” reaction mechanisms

for phytoene desaturation.

(A) The active center of OsPDS-His6 is depicted with the substrate cavity surface

given as grey shadow and FAD as orange sticks. All charged residues within 8 Å of

the redox-active isoalloxazine moiety are given as sticks color-coded by elements

(red, oxygen; blue, nitrogen; grey, carbon). The redox-active N1-N5 functionalities of

the isoalloxazine are postulated as the sole catalysts of carotene desaturation and

isomerization in PDS. (B) “Flavin only” reaction mechanism involving carotene radical

formation. A complete transfer of redox equivalents yields a reduced flavin and a

trans double bond at C11-C12. Reverse electron transfer from the reduced flavin to

the double bond yields a carotene radical that is stabilized by delocalization

allowing rotation (trans-cis isomerization) at C9-C10. (C) “Flavin only” reaction

mechanism involving carotene carbocation formation. The flavin mediates hydride

abstraction, yielding a FAD anion and a delocalization-stabilized carbocation that

allows rotation of the molecule at C9-C10. The FAD anion catalyzes proton

abstraction leading to formation of a trans double bond at C11-C12. The sites of

desaturation (C11-C12) and isomerization (C9-C10) are given in red. Fl denotes FAD.

B

C

2

Glu451

His159

Arg300

Glu542

Lys142

A

Page 73: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

59

Fig. 4-6 LC-MS analysis of ζ-carotene formed in buffer containing heavy water.

(A) The formation of ζ-carotene was detected photometrically at 400 nm upon

incubation in buffers with heavy water (D2O) and H2O. The UV/VIS spectrum of the ζ-

carotene formed is given as inset. (B) XIC for [M+H]+ of ζ-carotene (calculated m/z =

541.47678 ± 5 ppm) verifying its formation in the absence and the presence of D2O

and indicating that no 2D was incorporated into the carotene. (C) The

corresponding MS1 spectra of ζ-carotene formed with H2O (left) and D2O (right) in

buffers reveal identical [M+H]+ and M+1 signals, excluding deuterium incorporation.

4.1.4 Kinetic mechanism of the bi-substrate reaction in OsPDS-His6

As detailed in 1.3.1.1, PDS catalyzes the desaturation of 15-cis-phytoene in a

bi-substrate reaction also requiring plastoquinone (PQ) as a directly

interacting co-substrate. PQ acts as diffusible terminal electron acceptor to

reoxidize the enzyme-bound FADred formed upon carotene desaturation

(Gemmecker et al., 2015). The PDS structure suggests that phytoene and PQ

occupy the same cavity but cannot bind simultaneously, both fully

occupying the cavity. This suggests an ordered ping pong bi bi mechanism in

which the two substrates bind in sequential order. Moreover, this implies that

min min12 13 14 15 16

14.13

D2OH2O

14.15

Rel

ativ

e ab

und

ance

C40H61

m/z = 541.47407- 541.47949

A B

C

535 536 537 538 539 540 541 542 543m/z

0

10

20

30

40

50

60

70

80

90

100

Rel

ativ

e A

bun

dan

ce

541.47664

542.47992

PDS + H2O[M

+H

]+

14.05-14.20 min

535 536 537 538 539 540 541 542 543m/z

541.47661

542.47989

PDS + D2O

0

10

20

30

40

50

60

70

80

90

100

Rel

ativ

e A

bun

dan

ce

14.05-14.20 min

399 nm

297

12 13 14 15 16

14.09

mA

U(4

00 n

m)

D2O

14.13

H2O

424379

M+1

[M+

H]+

M+1

Page 74: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

60

carotene desaturation can per se occur in the absence of PQ, i.e. carotene

desaturation and PQ reduction are two successive, independent reactions.

Consequently, OsPDS-His6 should be able to desaturate phytoene in the

absence of DPQ. In fact, when DPQ was omitted in PDS assays, 1.25 nmol of

flavinylated PDS monomers led to the formation of 0.66 ± 0.01 nmol

phytofluene and 0.61 ± 0.14 nmol ζ-carotene within 30 min (n = 2). This equals

1.88 nmol of formed double bonds which is not far off the protein

concentration used. Consequently, each OsPDS-His6 monomer would

introduce one single double bond in the absence of DPQ. These findings

support that phytoene and phytofluene desaturation per se is

thermodynamically favored and independent of PQ. PQ is only required to

reoxidize the enzyme-bound FADred and to enable repeated cycles of

carotene desaturation. Consequently, PQ supply enables the progression of

phytoene desaturation and controls it kinetically. Additionally, despite the

presence of oxygen, no carotene desaturation beyond equimolarity (see

above) was observed. This supports that oxygen cannot serve as an electron

acceptor (Gemmecker et al., 2015). This is in contrast to CrtI in which FAD

reoxidation is dependent on molecular oxygen (Schaub et al., 2012).

The ordered ping pong bi bi mechanism is further supported by the fact that

PDS catalyzes the desaturation of 15-cis-phytoene to 9,15,9’-tri-cis-ζ-carotene

via the intermediate 9,15-di-cis-phytofluene (see 1.3.1.1), the latter being

released in substantial amounts (Gemmecker et al., 2015). This pleads for the

independency of the two formally identical carotene desaturation reactions

that take place at phytoene half sides. The intermediate is released from the

enzyme before being rebound by its undesaturated half to be converted into

ζ-carotene. The structure of OsPDS-His6 supports this notion because every

monomeric PDS subunit possesses one FAD to catalyze a single carotene

desaturation reaction. The intermediate needs to be expelled into the

membrane to allow reoxidization by PQ (see above). This suggests that each

monomeric subunit within the PDS homotetramers can be independently

Page 75: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

61

active. This raises questions on the mechanistic relevance of homotetrameric

assembly of OsDPS-His6 at membrane surfaces (see 4.1.1.2).

Taken together, the data presented support that PDS employs the following

ordered ping pong bi bi mechanism: Phytoene is bound in the substrate

cavity and converted to phytofluene that is then released. PQ now occupies

the same cavity to reoxidize the FADred formed upon phytoene desaturation.

In a subsequent round of carotene desaturation, PDS then binds either

phytoene or the saturated half of phytofluene as substrate. Accordingly, PDS

represents a bifunctional phytoene and phytofluene desaturase, with both

carotenes competing for binding. The two-step desaturation reaction from

phytoene to ζ-carotene resembles a two-enzyme cascade with the

intermediate phytofluene being in a steady state (McClure et al., 1969). This is

to be considered upon kinetic characterization (see 4.2).

4.2 Kinetic characterization of PDS and mathematical modeling

The investigation of the kinetics of OsPDS-His6 in the established biphasic

liposomal assay has been defined as a main aim of this thesis. This should

allow scrutinizing the functional implications deduced from the enzyme

structure. In this context, the mechanistic relevance of homotetrameric

assemblies in the active state of the enzyme at membrane surfaces (see

4.1.1.2) and in crystallo is of special interest (see 4.1.4). Traditional kinetic

investigations complemented by mathematical modeling (in collaboration

with the team of Prof. Dr. Timmer, Department of Physics, University of

Freiburg) were employed towards a better understanding of the sub-

processes involved in PDS catalysis.

4.2.1 Basic characterization of the OsPDS-His6 reaction

Using the biphasic liposomal assay established for OsPDS-His6 by Gemmecker

et al. (2015), the dependency of the OsPDS-His6 reaction on the added

Page 76: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

62

enzyme amount, pH and temperature was investigated. An approximately

linear correlation with ζ-carotene formation was observed for up to ca. 100 µg

PDS per assay (Fig. 4-7 A)), with end product formation then being limited by

phytoene supply (> 80 % conversion). Phytofluene levels remained unaffected

at increasing OsPDS-His6 amounts. This is consistent with this intermediate

being in a steady state (McClure et al., 1969; see 4.1.4). Consequently,

increasing enzyme amounts favored only ζ-carotene formation. Regarding

temperature and the pH, phytofluene formation again remained fairly

unaffected presumably due to a steady state situation whereas ζ-carotene

formation showed a broad pH optimum at pH ≈ 6.0 (Fig. 4-7 B) and

temperature optimum close to 40 °C (Fig. 4-7 C). These optima defined the

standard conditions for OsPDS-His6 assays at a saturating plastoquinone

concentration (see 4.2.3; Fig. 4-7 legend).

Reaction time courses of the OsPDS-His6 reaction (Fig. 4-8 D) revealed a rapid

decrease in phytoene within the first 10 min of incubation, accompanied by

rapid formation of 9,15-di-cis-phytofluene and 9,15,9’-tri-cis-ζ-carotene. As is

typical of a two-step reaction, the intermediate phytofluene approached a

steady state situation after 5 min and formation of the end product ζ-

carotene set in with a lag phase as it depends on the availability of the

intermediate (McClure et al., 1969). Strikingly, after 30 min phytoene and

phytofluene conversion ceased although 30 % of phytoene were left

unconverted.

An approach to equilibrium attaining 50 % ζ-carotene might be an

explanation. Alternatively, OsPDS-His6 might undergo inactivation, as

frequently observed with purified enzymes (Copeland, 2000). In fact, the

addition of fresh enzyme during the plateau phase restarted phytoene

conversion, indicating that PDS inactivation had occurred (Fig. 4-7 D), and led

to an equilibrium of > 95 % ζ-carotene. In contrast, no additional ζ-carotene

was formed upon disturbing a potential equilibrium by the addition of

substrates (phytoene or DPQ) during the plateau phase. Moreover, the

Page 77: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

63

reverse reaction carried out with ζ-carotene and phytofluene in the presence

of DPQH2, produced by DPQ reduction with diaphorase (Yu et al., 2014), did

not result in detectable formation of saturated carotenes. In summary, this

supports that reverse reactions of the catalytic steps in PDS are negligibly slow

and that the equilibrium is far on the side of the end product.

Fig. 4-7 Basic characterization of the OsPDS-His6 reaction.

(A – C) protein, pH and temperature dependency of phytofluene and ζ-carotene

formation from phytoene. Circle, phytofluene; square, ζ-carotene. Data represent

the mean of duplicates (A, C) or triplicates (B) ± SEM. Data were fitted with a spline in

A and C and with the dibasic pH equation in B. Each experiment was carried out

using the optimum values of the respective non-variable parameters, e.g. pH 6.0, 37

°C in A, etc. These optimum values determined the standard conditions: 25 µg PDS

per assay, 10 mM phytoene, (≈ 0.2 x KM; see 4.2.3) 19.25 mM DPQ (≈ 14 x KM; see

4.2.3), pH 6.0, 37 °C at 10 min incubation. (D) Reaction time course of phytofluene

and ζ-carotene formation from phytoene. Triangle, phytoene; circle, phytofluene;

0 50 100 1500.0

0.1

0.2

0.3A B

25 30 35 40 45 500

2

4

6

8

Temp [°C]

v [n

mol

min

-1m

g-1]

pK=4.5

pK=8.5

C

v [n

mol

min

-1m

g-1]

0 10 20 30 40 50 600

1

2

3

4

5

6 + PDS

D

6.0

5.0

4.0

3.0

2.0

1.0

0.0

pHPDS [µg]

v [n

mol

min

-1m

g-1]

4 5 6 7 8 9

1

6

5

2

3

4

Car

oten

es[n

mol

]

Time [min]

Page 78: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

64

square, ζ-carotene. Data represent single measurements. Asterisks denote the

activation of phytofluene and ζ-carotene formation upon the addition of 25 µg of

fresh PDS during the plateau phase after 30 min. Data were fitted with a spline.

4.2.2 Mathematical modeling of OsPDS-His6 reaction time courses

As laid out in 4.1.4, PDS employs an ordered ping pong bi bi mechanism

comprising three per se independent reactions – phytoene desaturation,

phytofluene desaturation and plastoquinone reduction. Every monomeric

subunit within a PDS homotetramer can act independently despite the

observation that homotetramers are the minimal assembly required for

enzymatic activity and represent the active state at membrane surfaces (see

4.1.1.2). Thus, the relevance of these oligomers was doubtful.

Traditional Michaelis-Menten kinetics cannot capture and explain kinetic

phenomena encountered with homooligomeric enzymes. Therefore, a

mathematical model of the PDS reaction based on mass action kinetics was

envisioned that could describe OsPDS-His6 reaction time courses as well as

substrate concentration dependencies. Mathematical modeling of reaction

time courses would allow determining rate constants for the main processes,

enable a better understanding of the enzyme kinetics and potentially provide

a testable hypothesis on the functional relevance of homomers.

As detailed above, a model of the PDS overall reaction (Fig. 4-9) necessarily

comprises three fundamental, per se independent processes:

(i) the desaturation of phytoene (p) to phytofluene (pf)

(ii) the desaturation of phytofluene (pf) to ζ-carotene (z)

(iii) reoxidation of the intermediary electron acceptor FADred by

reduction of the oxidized terminal acceptor DPQ (Q)

Additionally, as witnessed by rapid loss of activity in vitro (see 4.2.1), a fourth

process needs to be considered for modeling of reaction time courses:

(iv) time-dependent enzyme inactivation

As established in Michaelis-Menten kinetics, the three catalytic main

processes consist at minimum of three sub-processes that represent equilibria

Page 79: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

65

of forward and reverse reactions: association and dissociation of enzyme and

substrate, desaturation and saturation of substrate as the catalytic step, as

well as dissociation and association of enzyme and product. The “monomeric

model” for the PDS reaction (Fig. 4-8 A) assumes, as laid out in 4.1.4, that each

monomer within the homotetramer acts independently regarding the main

catalytic processes (i – iii). Thus, intermediary phytofluene pf is expelled from

PDS into the hydrophobic liposome core, diffuses amidst the phytoene

molecules deposited into liposomes as initial substrate and will eventually be

bound and converted by an oxidized PDS subunit (the same one after

reoxidation by DPQ or a different one).

Alternatively, taking homotetrameric assemblies into account, the following

“substrate channeling model” is conceivable (Fig. 4-8 B): In crystallo, the

substrate channels of the homotetramer subunits point towards each other

with only few Å between the entrances (Brausemann et al., 2017). Thus, as

phytoene and phytofluene desaturation are formally identical reactions on

substrate half sides, one can intuitively assume that intermediary phytofluene

formed by one subunit can be directly channeled into an adjacent oxidized

one of the same homotetramer. This means that upon release from the first

subunit, this phytofluene population – termed nascent phytofluene pf* – is

restricted in diffusion occupying a membrane microdomain below the

homotetramer. Two fates of phytofluene pf* may coexist. The non-channeled

fate corresponds to the situation in the “monomeric model” in which pf* is in

any case released into the hydrophobic membrane core and then represents

freely diffusing pf (right side Fig. 4-9 B). In the channeled fate, pf* can be

rapidly channeled into another oxidized subunit within the same

homotetramer by substrate channeling to be more directly converted to ζ-

carotene (right side Fig. 4-8 B). Substrate channeling as a form of cooperation

between enzyme subunits requires various substrate cavities within short

distances to restrict diffusion (Wheeldon et al., 2016), a prerequisite that is met

by PDS homooligomers. It would facilitate and accelerate end product

formation and progression of the entire carotene desaturation pathway.

Page 80: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

66

Fig. 4-8 Kinetic schemes and models of the PDS overall reaction.

(A) Monomeric model: PDS monomeric subunits (rectangles) within the

homotetramer work independently. Orange/blue color denotes reduced/oxidized

half sides of phytoene (p), phytofluene (pf) and ζ-carotene (z) and the respective

redox state of the PDS-bound FAD. The overall reaction comprises the three main

processes phytoene desaturation (i), phytofluene desaturation (ii) and plastoquinone

reduction (iii) with the rate constants kp, kpf and krox, respectively. Each rate constant

encompasses three equilibria represented by the reaction arrows associated to

each of the three main processes as highlighted by shadowed areas: association-

dissociation of enzyme and substrate, desaturation-saturation of substrate and

dissociation-association of enzyme and product. All hydrophobic carotene

substrates and decylplastoquinone (Q) are soluble in the hydrophobic core of

liposomal membranes. (B) Substrate channeling model: It accounts for substrate

channeling between subunits within a homotetramer. Two populations and fates of

Membrane

QH2

kpf

kp

krox

(i)

(ii)

(iii)

A

Membrane

p

pf

z

Q

QH2

kdif

kp

kpf

kp

krox

(i)

(ii)

(iii)

B

kpf*

Q

kage

z

pf

pf

p

QH2

kage

kage kage

p p

pf*

pf* pf

z

zz

Q

QH2

pf

pf

p

Q

zpf *

pf*

pf*

Monomeric model Substrate channeling model

(iv)

(iv)

Page 81: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

67

phytofluene coexist. Left; nascent phytofluene (pf*) that is produced from phytoene

(p) can be restricted in its diffusion into the membrane upon release from the

enzyme, thus remaining in proximity to the homotetramer as indicated by the bent

arrow. It can be channeled into a second subunit of the same homotetramer

containing FADox, allowing rapid conversion to ζ-carotene (z) with the rate constant

kpf*. Right; pf* can alternatively diffuse into PDS-distant membrane areas with rate

constant kdif, then defining the population pf. From there it can be taken up and

converted into ζ-carotene (z) with rate constant kpf. The process kage, represents

enzyme inactivation (iv) which refers to both the reduced and oxidized state of PDS.

Both models were implemented into mathematical models, each

represented by a set of ordinary differential equations (ODEs). Based on mass

action kinetics, these allow simulating and describing the changes of all

reactants – namely p, pf, z, Q and FAD (representing the catalytic

components of each PDS subunit) – during PDS reaction time courses by

providing the initial amounts of reactants at t = 0 min (Table 3).

Table 3 Initial amounts of reactants in reaction time course with OsPDS-His6.

Reactant amounts in the OsPDS-His6 reaction time course assays are given in [nmol]

for t = 0 min. Q * indicates that the amount of oxidized quinone was held constant

throughout reaction time courses in modeling because rapid non-enzymatic

reoxidation of DPQH2 in liposomes was observed (e.g. despite 2 e- transfer involved in

both carotene desaturation and DPQ reduction, up to 26 carotene desaturation

reactions per DPQ were observed experimentally).

reaction time

course

FADox

[nmol]

FADred

[nmol]

p

[nmol]

pf

[nmol]

z

[nmol]

oxidized Q

[nmol]

p high 0.18 0 3.7 0 0 9.6 *

p low 0.18 0 1.3 0 0 9.6 *

pf 0.18 0 0 5.2 0 9.6 *

A rate constant k (parameter) is assigned to every sub-process, reflecting its

reaction velocity. Both models should be evaluated in order to distinguish

which would best describe PDS behavior in three reaction time courses that

Page 82: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

68

were conducted for this purpose. Two contained different initial phytoene

amounts (p high, p low) and one contained phytofluene as a substrate (pf)

(Table 3). The aim was to determine which model could provide a single set

of rate constants to adequately describe all reaction time courses

(simultaneous parameter estimation; opposed by individual parameter

estimation, providing a set of rate constants for every single reaction time

course). Mathematical modeling was initiated with the “monomeric model”

including all 18 forward and reverse reactions but no enzyme inactivation

(Fig. 4-9).

Fig. 4-9 “Monomeric PDS model” including all forward and reverse reactions.

For details, see legend Fig. 4-9 A. All forward and reverse reactions of sub-processes

are considered and were assigned a rate constant (R indicates reverse reactions).

The following set of ODEs and rate constants (parameters) was used to

describe the amounts of p, pf, z, Q and oxidized and reduced FAD over time

(“o” indicates complexation; R indicates reverse reaction):

p

kr2

kr1

kr3

Me

mb

rane

pf

p

pf

pf pf

zz

QQ

QH2QH2

kRp

kRr2

kRr3

kf2

kRf2

kRf

kf

kRz

kz

kRpf

kpf

kpf2

kRpf2

kp

kRr1

(i)

(ii)

(iii)

Page 83: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

The large number of reversible reactions

plateaus. Thus, it describe

individual parameter estimation

enzyme inactivation (as the

However, the model fail

simultaneous parameter estimation even

reaction time courses “p high” and “p low”

Fig. 4-10 Observed data and modeling of

model” including forward and reverse reactions

The observables p (black), pf (red) and z (blue)

high” (A) and “p low” (B)

Time [min]

p high

Car

oten

es[n

mol

]

A

Results

large number of reversible reactions allows steady states, i.e. modeling of

described the plateaus of each reaction time course

estimation (dashed lines Fig. 4-10), without implementing

as the proven cause for observed plateaus

model failed to describe the reaction time course

simultaneous parameter estimation even when only the

p high” and “p low” were included (Fig. 4

Observed data and modeling of reaction time courses with the “monomeric

model” including forward and reverse reactions.

p (black), pf (red) and z (blue) during the reaction time

(B) are given as data points, representing the mean

Time [min]

B

Car

oten

es[n

mol

]

p low

pf

z

p

pf

z

p

69

, i.e. modeling of

reaction time course upon

without implementing

observed plateaus; see 4.2.1).

reaction time courses upon

two phytoene

Fig. 4-10).

with the “monomeric

reaction time courses “p

are given as data points, representing the mean of

Page 84: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

70

triplicates ± SEM. The predictions are based on the “monomeric model” including all

forward and reverse reactions (Fig. 4-9) and are given as lines, either upon individual

parameter estimation (dashed) or upon simultaneous parameter estimation (solid).

The model proved to be overparameterized (Chen et al., 2010): Estimated

parameter values were poorly determined and could vary over a wide range

without impairing the goodness of fit (parameter likelihood profiles not

shown). This is encountered when model complexity (the degrees of freedom)

is too high in relation to the available data describing the modeled system.

This is the case with 18 sub-processes but only three time-resolved observables

(p, pf and z). To reduce model complexity and tailor it to the available data,

only the most relevant processes were included. Successive rounds of model

reduction and reevaluation indicated the feasibility of narrowing down the

model to the three main catalytic processes (i – iii) with one rate constant

each – namely phytoene desaturation (kp), phytofluene desaturation (kpf)

and plastoquinone reduction (krox) (Fig. 4-8 A, shaded areas). Moreover,

enzyme inactivation (iv; implemented by decreasing the amount of oxidized

and reduced FAD over time with rate constant kage) was required as an

essential process to describe the plateaus because no reverse reactions were

considered, as they were found to be negligibly slow (see 4.2.1). The simplified

“monomeric model” comprised the four rate constants in a set of five ODEs:

���p� = − ∙ �p� ∙ �FAD��� (1)

���pf� = −� ∙ �pf� ∙ �FAD��� + ∙ �p� ∙ �FAD��� (2)

���z� = � ∙ �pf� ∙ �FAD��� (3)

���FAD��� = − ∙ �p� ∙ �FAD��� − � ∙ �pf� ∙ �FAD��� + ��� ∙ �Q� ∙ �FAD���� − ��� ∙ �FAD��� (4)

���FAD���� = ∙ �p� ∙ �FAD��� + � ∙ �pf� ∙ �FAD��� − ��� ∙ �Q� ∙ �FAD���� − ��� ∙ �FAD���� (5)

This allowed describing the three reaction time courses, however, only upon

individual parameter estimation (not shown). Upon simultaneous parameter

estimation z formation was not adequately described in “p low” and “p high”

Page 85: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

71

(solid lines in Fig. 4-11 A, B) while pf formation was generally well described –

and so was z formation for “pf” (solid lines in Fig. 4-11 C). Subsequent

evaluation revealed that individual estimation of kpf for every reaction time

course was sufficient to fit the data, while all other parameter values could be

estimated simultaneously (dotted lines in Fig. 4). The deduced rate constants,

with the varying kpf values for the three reaction time courses, are summarized

in Table 2. The difference between kpf in the “p low” and “p high” reaction

time courses is insignificant, both being ca. 5 nmol-1 min-1. In contrast, kpf for

the “pf” reaction time course is as low as 1.1 ± 0.1 nmol-1 min-1. Consequently,

the model suggests that the conversion of “nascent” pf produced from

phytoene (as in “p high” and “p low) proceeds 5 x faster at the same

concentration of reactants than the conversion of pf experimentally

deposited in liposomes (as in “pf”).

Fig. 4-11 Observed data and modeling of reaction time courses with the “monomeric

model”.

0 10 20 30 40 50 60

0.5

1.0

1.5

2.0

2.5

3.0

3.5

0 10 20 30 40 50 600.0

1.0

2.0

3.0

4.0

5.0

0 10 20 30 40 50 600.0

0.2

0.4

0.6

0.8

1.0

1.2

p

pf

z

z

pf

p

Time [min] Time [min]

Time [min]

pf

z

p

pf

p high p low

Car

oten

es[n

mol

]

Car

oten

es[n

mol

]

Car

oten

es[n

mol

]

B

C

A

Page 86: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

72

The observables p (black), pf (red) and z (blue) during the reaction time courses “p

high” (A) and “p low” (B) are given as data points, representing the mean of

triplicates. The model predictions are based on the “monomeric model” (Fig. 4-8 A)

only comprising kp, kpf, krox and kage, either upon simulatenous parameter estimation

for all three reaction time courses (solid lines) or upon simultaneous parameter

estimation for kp, krox and kage and individual estimation for kpf (dashed lines).

Apparently, the model requires two kinetically inequivalent phytofluene

populations for simultaneous parameter estimation. These requirements are

met by the “substrate channeling model” introduced in Fig. 4-8 B, allowing

two fates of “nascent” intermediary phytofluene pf*, namely free diffusion in

the membrane core and restricted diffusion in vicinity of the PDS tetramer.

The “substrate channeling model” was implemented with the following ODEs:

���p� = − ∙ �p� ∙ �FAD��� (1)

���pf ∗� = −�∗ ∙ �pf

∗� ∙ �FAD��� −��� ∙ �pf∗� + � ∙ �p� ∙ �FAD��� (6)

���pf� = −� ∙ �pf� ∙ �FAD��� + ��� ∙ �pf

∗� (7)

���z� = � ∙ �pf� ∙ �FAD��� + �∗ ∙ �pf

∗� ∙ �FAD��� (8)

���FAD� = − ∙ �p� ∙ �FAD��� − � ∙ �pf� ∙ �FAD��� − �∗ ∙ �pf

∗� ∙ �FAD��� + ��� ∙ �DPQ� ∙

�FAD���� − ��� ∙ �FAD��� (9)

���FAD���� = ∙ �p� ∙ �FAD��� + � ∙ �pf� ∙ �FAD��� + �∗ ∙ �pf

∗� ∙ �FAD��� − ��� ∙ �DPQ� ∙

�FAD���� − ��� ∙ �FAD���� (10)

In fact, the “substrate channeling model” fitted all reaction time courses well

upon simultaneous parameter estimation (Fig. 4-12 A – C). The resulting set of

rate constants is given in Table 4. The parameter likelihood profiles (Fig. 4-13)

demonstrate that all parameters are well defined, i.e. their values cannot vary

without impairing the goodness of fit.

Page 87: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

73

Fig. 4-12 Observed data and modeling of the PDS reaction time courses with the

“substrate channeling model”.

(A – C) The observables p (black), pf (red) and z (blue) during the reaction time

courses “p low” (A), “p high” (B) and “pf”(C) are given as data points representing

the mean of triplicates. The model prediction (lines) is based on the “substrate

channeling model” (equ. 1 and 6 – 10; Fig. 4-8 B) and simultaneous parameter

estimation. Shadowed areas indicate one standard deviation as estimated by the

error model. (D) Prediction of the amount of oxidized, active (ox) and reduced (red)

PDS upon simultaneous parameter estimation. (E, F) Predicted carotene fluxes in the

reaction time course “p high” through the different sub-processes of the model,

denoted by their rate constants (Fig. 4-8 B). Note the different scaling in E and F.

0 10 20 30 40 50 600.0

0.2

0.4

0.6

0.8

1.0

1.2

1.4

0 10 20 30 40 50 600.0

1.0

2.0

3.0

4.0

5.0

0 10 20 30 40 50 600.0

0.1

0.2

0.3

0.4

kpkpf*

krox

Car

oten

es[n

mol

]

Time [min]

Time [min]

B

C D

FA

D [n

mol

]

0 10 20 30 40 50 600.0

0.1

0.2

Time [min]

pf

E F

ox

red

Car

oten

es[n

mol

min

-1]

Time [min]

p high

Time [min]

Car

oten

es[n

mol

min

-1]

pf

z

0 10 20 30 40 50 600.000

0.001

0.002

0.003

0.004

0.005 p highkdif

kpf

Time [min]

A

0 10 20 30 40 50 60

0.5

1.0

1.5

2.0

2.5

3.0

3.5

pf

z

p

Car

oten

es[n

mol

]

Car

oten

es[n

mol

]

p high

pf

z

p

p low

Page 88: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

74

Fig. 4-13 Likelihood profiles for the parameters values deduced from the “substrate

channeling model”.

The profile likelihood, ∆χ2, is plotted over a range of parameter values around the

estimated optimal value marked by a dot. As reference, the 68 % / 90 % / 95 %

confidence level (CL) thresholds corresponding to ∆χ2 = 1 / 2.71 / 3.84 are given as

horizontal lines.

Table 4 Parameter values deduced from the “monomeric model” and the “substrate

channeling model”.

Parameter values deduced from the PDS reaction time courses “p high”, “p low”

and “pf” with the “monomeric model “(Fig. 4-8 A) and “substrate channeling model”

(Fig. 4-8 B) are given. Parameter values are given as mean ± SD. For the “monomeric

model”, simultaneous parameter estimation was applied to krox and kage, assuming

that FAD reoxidation and enzyme inactivation are independent of the carotene

substrate present (p or pf), and to kp. Individual parameter estimation for every

reaction time course was applied to kpf. For the “substrate channeling model”,

kp kpf kdif

kpf* krox kage

Parameter value

Page 89: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

75

simultaneous parameter estimation was applied to all parameters across all reaction

time courses.

Monomeric model

Parameter Value

Substrate channeling model

Parameter Value

Phytoene

conversion

�� 0.54 ± 0.02 nmol-1 min-1 �� 0.55 ± 0.02 nmol-1 min-1

Phytofluene

conversion

�� (pf) 1.14 ± 0.04 nmol-1 min-1 �� 1.15 ± 0.04 nmol-1 min-1

�� (p high) 5.10 ± 0.24 nmol-1 min-1

�� (p low) 4.77 ± 0.22 nmol-1 min-1

- - �� ∗ 5.44 ± 0.32 nmol-1 min-1

- - �!" 0.02 ± 0.01 min-1

PDS

reoxidation

�#$% 5.76 (-1.92 +5.84)

nmol-1 min-1

�#$% 5.40 ( − 1.86 + 5.67)

nmol-1 min-1

PDS

inactivation

�&'(

0.22 ± 0.01 min-1

�&'( 0.22 ± 0.01 min-1

According to the rate constants, the conversion of lipid-diffusible p (kp ≈ 0.55

nmol-1 min-1) is only half as fast as the conversion of lipid-diffusible pf (kpf ≈ 1.15

nmol-1 min-1). However, both substrates in the hydrophobic liposome core are

converted more slowly than “nascent” pf* (kpf* ≈ 5.44 nmol-1 min-1). According

to the working hypothesis, the restricted diffusion of pf* in a membrane

microdomain close to the PDS homotetramer increases its local

concentration and allows a different subunit of the same PDS homotetramer

to rapidly bind and convert it into z. Thus, conversion of pf* is 5 x faster than

conversion of pf (at the same reactant concentrations). It is worth noting that

the reoxidation of FADred by Q is very fast with a krox of 5.40 nmol-1 min-1 (up to

11.17 nmol-1 min-1 within one standard deviation). As a result, the PDS

reoxidation flux through krox is high and easily compensates for the fluxes

through PDS-reducing carotene desaturation processes through kp, kpf and

kpf*, (Fig. 4-12 E). Accordingly, only a small proportion of PDS is predicted to be

Page 90: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

76

in its reduced state during reaction time courses (Fig. 4-12 D). This suggests

that PDS reoxidation is not rate-limiting. Regarding inactivation, the model

predicts decreasing levels of the oxidized and reduced OsPDS-His6 with kage ≈

0.22 min-1, meaning that 22 % are inactivated each minute and resulting in a

half life of ca. 4 min (Fig. 4-12 D). The rate constant kdif (≈ 0.02 min-1),

representing the release of nascent pf* from the membrane microdomain

into the hydrophobic membrane core as freely diffusing pf, suggests that only

2 % of pf* leave the microdaim each minute. Accordingly, the calculated

carotene fluxes through all carotene desaturation processes (Fig. 4-12 E) show

that the pf* flux into z through kpf* exceeds by far the phytofluene fluxes

through kdif and kpf. Thus, pf* has a low probability of escaping channeling

that strongly accelerates end product formation.

In summary, the “substrate channeling model” – in contrast to the

“monomeric model” – describes all PDS reaction time courses well with a

single set of rate constants. Thus, substrate channeling appears as necessary

process to describe PDS behavior. This finding corroborates the relevance of

PDS homotetramers in the enzyme’s active state at membrane surfaces (see

4.1.1.2). The model suggests that channeling of phytofluene between the

subunits, i.e. restricting intermediate diffusion, allows more rapid formation of

the end product ζ-carotene than possible with monomeric action of PDS.

4.2.3 Substrate concentration-dependent kinetics of OsPDS-His6

PDS catalyzes a bi-substrate reaction involving a carotene, either phytoene or

phytofluene, and DPQ. In order to investigate the concentration-dependent

enzyme behavior by their Michaelis-Menten parameters, pseudo-first order

conditions were attained in assays. The concentration of the carotene

substrate was varied while using the quinone substrate at invariable,

saturating concentrations – and vice versa. Additionally, the validity of the

“substrate channeling model” was tested by investigating whether it would

Page 91: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

77

adequately simulate substrate concentration dependencies based on the

rate constants (Table 4) and initial enzyme and substrate amounts (Table 3).

First, the concentration dependency on DPQ was examined at the maximally

attainable concentration of 40 mM phytoene (higher concentrations lead to

liposome precipitation). The formation of ζ-carotene could be fitted with the

Michaelis-Menten (MM) equation (Fig. 4-14 A). Since the intermediate

phytofluene is in a steady state, its formation did not show MM conformity.

Consequently, product:intermediate ratios varied substantially (dotted line in

Fig. 4-15 A), with increasing DPQ concentrations favoring end product

formation by OsPDS-His6. Simulation of the DPQ dependency with the

“substrate channeling model” revealed the same trend (compare Fig. 4-15 A

and D). While the observed and estimated apparent Vmax values were very

similar, there was a ca. 4-fold difference in the respective KM values (Table 5).

Second, the enzyme’s dependency on phytoene and phytofluene

concentration was examined under DPQ saturation (19.25 mM ≈ 15 x KM; Fig.

4-15 B and C). Both carotene substrates could not be increased to saturation

for reasons of liposome integrity (see above). Fitting ζ-carotene formation

from phytoene with the MM equation (Fig. 4-14 B) allowed determining

apparent phytoene KM and Vmax values that are in agreement with the

simulated values (Table 5). Again, the formation of the intermediate did not

show MM conformity. Notably, no sigmoidality – a hallmark of cooperative

substrate binding in homooligomeric enzymes – was observed.

As seen with DPQ, the inequivalent kinetic behavior of phytofluene and ζ-

carotene formation affects the product:intermediate ratio in a substrate

concentration-dependent manner. Increasing phytoene concentrations

favored phytofluene release with the z:pf ratio decreasing from ca. 4:1 to 1:1

(Fig. 4-14 B, dotted line). These relations are well reflected in the simulation of

the phytoene dependency (compare Fig. 4-14 B and E). Thus, low carotenoid

fluxes through PDS favor end product formation, while the opposite favors

intermediate release. For the phytofluene dependency, ζ-carotene formation

Page 92: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

78

could be fitted with the Michaelis-Menten equation (Fig. 4-14 C), revealing

apparent KM and Vmax values that differ from those derived from simulation

(compare Fig. 4-14 C and F; Table 5).

Fig. 4-14 Data and model predictions for substrate concentration-dependent kinetics

of OsPDS-His6.

Measured (A-C) and simulated (D-E) concentration dependency of the PDS

reaction. (A) DPQ concentration dependency determined at a fixed 40 mM (≈ 1 x

KM) phytoene concentration. (B) Phytoene concentration dependency measured at

a fixed 19.25 mM (≈ 15 x KM) DPQ concentration. (C) Phytofluene concentration

0 10 20 30 40 500

5

10

15

20

0 10 20 30 40

0

5

10

15

20

0 10 20 30 40 500

5

10

15

20

25

30

0 1 2 3 4 5 6 7 80

5

10

15

20

0 1 2 3 4 5 6 7 80

5

10

15

20

25

0

1

2

3

DPQ [mM]

v [n

mo

lmin

-1 m

g-1

]

0 10 20 30 400

5

10

15

20

0

2

4

6

8

Phytoene [mM]

B

Phytofluene [mM]

A

C

z/fp[m

mo

l/mm

ol]z/p

f[mm

ol/m

mol]v

[nm

olm

in-1

mg

-1]

v [n

mo

lmin

-1 m

g-1

]

v [n

mo

lmin

-1 m

g-1

]v

[nm

olm

in-1

mg

-1]

v [n

mo

lmin

-1 m

g-1

]

Phytoene [mM]

DPQ [mM]

D

E

Phytofluene [mM]

F

Page 93: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

79

dependency measured at a fixed 19.25 mM DPQ concentration. Data represent the

mean of triplicates ± SEM. Phytofluene and ζ-carotene formation in A – C were fitted

with the Michaelis-Menten equation (solid lines; goodness of fit for ζ-carotene

formation: A, R2 = 0.98; B, R2 = 0.97; C, R2 = 0.98). Phytofluene formation in B was fitted

with a spline. The ζ-carotene:phytofluene ratios in A and B were derived from the

measured pf and z amounts, are given as dotted lines and plotted to the right y-axis.

Red squares, ζ-carotene; blue squares = phytofluene. In D – F, the model prediction

of phytofluene and ζ-carotene formation is depicted as solid line (red, phytofluene;

blue, ζ-carotene). Shaded areas represent one standard deviation.

Table 5 Observed and estimated apparent KM and Vmax values for PDS substrates.

Apparent KM and Vmax values were determined based on ζ-carotene formation for

the observed and estimated concentration-dependent kinetics of OsPDS-His6 (Fig. 4-

14) and the mutated PDS enzyme Arg300Ser. Values represent the mean ± SD). The

observed data (obs.) were fitted with the Michaelis-Menten equation. Estimated

values (est.) were obtained by simulation with the “substrate channeling model”.

Substrate KM obs.

[mM]

KM est.

[mM]

Vmax obs.

[nmol min-1 mg-1]

Vmax est.

[nmol min-1 mg-1]

DPQ 1.3 ± 0.2 6.2 ± 0.2 28.1 ± 1.4 26.1 ± 0.6

Phytoene 53.9 ± 18.1 71 (-27 +160) 46.3 ± 10.7 51 (-18 +43)

Phytofluene 66.8 ± 20.7 126 (-40 +120) 48.4 ± 10.3 195 (-67 +187)

In summary and in support of the validity of the “substrate channeling model”,

the concentration-dependent relations of intermediate and end product

formation are well reflected across all simulations. However, the model

overestimates Michaelis-Menten parameters, used here as a quantitative

measure to allow comparisons, by factors of 1.1 to 4.1 (Table 5). This deviation

and its importance regarding model validity will be addressed in the

discussion (see 5.1.5).

Page 94: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

80

4.3 Investigations on herbicide resistance in PDS

As detailed in 1.3.1.3, phytoene desaturase PDS is a prominent target for a

structurally heterogenous group of bleaching herbicides, such as norflurazon

(NFZ), fluridone and diflufenican, that share a meta-trifluoromethylphenyl (m-

CF3-phenyl) moiety that is often fused to a second (heterocyclic) ring with a

carbonyl group. The lack of protein structural information has hampered the

molecular understanding of the mode of action of these compounds. The

refinement of the OsPDS-His6 structure in a complex with NFZ during this thesis

(in collaboration with the team of Prof. Dr. Einsle, Department of Chemistry,

University of Freiburg) has changed this situation allowing in depth

investigations. Such work is of interest in the light of emerging resistances

against commonly used herbicides in various crops.

4.3.1 Kinetic analysis of the mode of inhibition by norflurazon for OsPDS-His6

The OsPDS-His6 crystal structure shows NFZ bound to the assumed PQ binding

site within the substrate cavity. If true, NFZ and plastoquinone should compete

for binding to this site. Phytoene occupies the same substrate cavity (see

4.1.4) and therefore it might as well interfere with NFZ binding.

In order to investigate this, in vitro inhibition kinetics with norflurazon were

carried out with OsPDS-His6. NFZ inhibition could be overcome by increasing

DPQ concentrations, i.e. Vmax was attained in the presence of NFZ (Fig. 4-15

A). Thus, inhibition regarding PQ could be best described with the competitive

inhibition model, revealing a Ki of 0.23 ± 0.03 mM for NFZ. This is in accordance

with the previous finding that DPQ can displace NFZ during OsPDS-His6

purification (Gemmecker et al., 2015) and quinone-competitive inhibition

reported with cell-free in vitro assays by Breitenbach et al. (2001). Other meta-

trifluoromethylphenyl–containing PDS inhibitors such as fluridone and

diflufenican were also quinone-competitive (Fig. 4-15 C, D) and consequently

appear to also occupy the PQ binding site, as previously hypothesized by

Laber et al. (1999). In contrast, NFZ inhibition was not overcome by increasing

phytoene concentrations (Fig. 4-15 B), indicating that NFZ and phytoene do

Page 95: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

81

not compete and supporting either non-competitive or uncompetitive

inhibition by NFZ regarding the carotene substrate. Both inhibition models

described the inhibition kinetics poorly, e.g. revealing a Ki of 0.59 ± 0.09 mM

for non-competitive inhibition. Non-competitive inhibition is supported by the

fact that the crystallized PDS-NFZ complex represents an enzyme-inhibitor

complex formed in the absence of substrate, with uncompetitive inhibitors

only exhibiting affinity for the enzyme-substrate complex (Copeland, 2000).

Fig. 4-15 Mode of OsPDS-His6 inhibition by NFZ, fluridone and diflufenican

(A, B) Inhibition of OsPDS-His6 was investigated under increasing concentrations of

the inhibitor NFZ and of the substrates DPQ (A) or phytoene (B). Data represent

triplicates ± SEM and were fitted with the equations for competitive inhibition (A; R2 =

0.99) and non-competitive inhibition (B; R2 = 0.96). (C, D) Inhibition of OsPDS-His6 was

investigated under increasing concentrations of the substrate DPQ in the presence

of the inhibitors diflufenican (C) or fluridone (D). Data represent duplicates ± SEM and

were fitted with the equations for competitive inhibition (A, R2 = 0.96; B, R2 = 0.98).

DPQ [mM]

v [n

mol

ζ m

in-1

mg-1

]

0 5 10 15 200

1

2

3

4

5

6

0.0 mM0.6 mM1.2 mM2.4 mM

Phytoene [mM]

v[nm

olζ

min

-1m

g-1]

0 10 20 30 40 500

2

4

6

8

0.0 mM1.6 mM3.3 mM

A B

DPQ [mM]

v [n

mol

ζ m

in-1

mg-1

]

0 5 10 15 200

1

2

3

4

5

0.00 mM

0.16 mM

DPQ [mM]

v [n

mol

ζ m

in-1

mg-1

]

0 5 10 15 200

2

4

6

8

0.00 mM

0.27 mM

C D

FluridoneDiflufenican

Page 96: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

82

It is noteworthy that this mode of inhibition further corroborates the postulated

ordered ping pong bi bi mechanism of PDS because the observed

competition between NFZ and plastoquinone and the lack of competition

between NFZ and phytoene, all occupying the same substrate cavity, can

only be explained when sequential binding of the two substrates is warranted.

4.3.2 Norflurazon resistance conferred by point mutations

Point mutations of PDS conferring resistance to NFZ have been identified in

cyanobacteria, algae and an aquatic plant (see 1.3.1.3). Protein alignments

reveal that five highly conserved amino acids corresponding to Phe162, Arg300,

Leu421, Val505 and Leu538 in the O. sativa enzyme are affected (Fig. 4-16).

According to the structure of OsPDS-His6 in a complex with NFZ, these residues

are localized in the active center environment (Fig. 4-17).

Fig. 4-16 Protein alignment of PDS mutant enzymes conferring NFZ resistance.

The following residues are highlighted: 1, His159; 2, Phe162; 3, Met188; 4, Met277; 5, Ala280;

6, Arg300; 7, Met310; 8, Leu421; 9, Phe423; 10, Val505; 11, Tyr506; 12, Thr508 13, Leu538. Global

sequence alignment was carried out with the Blosum62 matrix. Identical residues are

green, similar residues greenish or yellow. Position numbering refers to the immature

10 13

8

6

1 3

4 7

11

9

5

2

12

Synechocystis

Oryza

Arabidopsis

Chlorella

Hydrilla

Synechococcus

Synechocystis

Oryza

Arabidopsis

Chlorella

Hydrilla

Synechococcus

Synechocystis

Oryza

Arabidopsis

Chlorella

Hydrilla

Synechococcus

Synechocystis

Oryza

Arabidopsis

Chlorella

Hydrilla

Synechococcus

Synechocystis

Oryza

Arabidopsis

Chlorella

Hydrilla

Synechococcus

Synechocystis

Page 97: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

83

protein from O. sativa (Acc. A2XDA1.2) including its N-terminal 87 amino acid transit

peptide. Organisms and accession numbers (from top to bottom): Oryza sativa,

A2XDA1.2; Arabidopsis thaliana, Q07356.1; Chlorella zofingiensis, ABR20878.1; Hydrilla

verticillata, AAT76434.1; Synechococcus elongatus PCC 7942, CAA39004.1;

Synechocystis sp. PCC6803, CAA44452.1.

Fig. 4-17 Conserved residues in OsPDS-His6 conferring NFZ resistance upon mutation

The inner surface of the substrate cavity is depicted. The substrate cavity entry in the

membrane binding domain is indicated by an arrow. The redox cofactor FAD is

given as sticks in orange. Conserved residues whose mutation has been reported to

convey NFZ resistance are given as sticks, color-coded by elements (grey, carbon;

blue, nitrogen; red, oxygen). The numbering relates to the Oryza sativa protein (Acc.

A2XDA1.2) including its N-terminal 87 amino acid transit peptide.

Their mutation might therefore have multiple effects. They might affect size

and general architecture of the active site, alter flavin binding or interfere

more specifically. The latter most likely applies to Arg300. It coordinates NFZ via

a hydrogen bond and – as indicated by competition and structural similarities

with the benzoquinone head group – it is assumed to be involved in the

binding of PQ (Brausemann et al., 2017). In order to investigate the role of

Arg300 in NFZ binding, the mutations Arg300Ser and Arg300Thr that were reported

to confer NFZ resistance (Arias et al., 2006) were introduced into OsPDS-His6.

Ser (C3) and Thr (C4) are several Å shorter than Arg (C7). Consequently, given

Page 98: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

84

the average distance of only about 3 Å between donor and acceptor of

hydrogen bonds (Kyte, 2006), Ser and Thr most likely cannot form hydrogen

bonds with NFZ despite their hydroxyl functionalities. Both purified PDS variants

retained ζ-carotene forming activity but at reduced levels with 15 % and 6 %

of the wild type enzyme, respectively. Being more active, Arg300Ser was

chosen for further characterization in comparison with wild type OsPDS-His6.

Flavinylation was not affected with both wild type OsPDS-His6 and Arg300Ser

PDS containing ca. 70 % holoenzyme. There was also no difference in

membrane association (Fig. 4-18 A). GPC analysis revealed that the mutation

did not affect protein solubility and homooligomeric assembly (Fig. 4-19 B).

Therefore, Arg300Ser PDS most likely maintained a native overall fold and

assembly and the loss of activity is likely caused by more subtle changes

affecting enzyme kinetics.

Fig. 4-18 Association with liposomal membranes and oligomeric assembly in solution

of wild type OsPDS-His6 and the Arg300Ser mutant enzyme.

(A) SDS-PAGE analysis (12 %, Coomassie-stained) of liposomal binding assays. Lanes

represent the liposome-bound PDS protein obtained from one PDS assay. WT, wild

type. (B) Elution traces of wild type OsPDS-His6 and the mutant enzyme Arg300Ser

monitored at 280 nm upon GPC (Superose 6 10/300 GL column). The dominant high

mass peak (oligo) represents the flavinylated and active homooligomer, the two

A B

5 10 15 20

5 10 15 200 ml

ml

VE

VE

A2

80

nm

Arg300Ser

OsPDS-His6

kDa

100

70

55

40

35

25

15

10

WT Arg300Ser

oligo

mono

FAD

oligo

mono

FAD

V0

V0

Page 99: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

85

lower mass peaks represent the unflavinylated, inactive monomer (mono) and free

FAD that has been released from the enzyme upon sample preparation and GPC.

In order to evaluate this hypothesis, the required kinetic data were gathered.

In fact, quinone-competitive inhibitions kinetics (Fig. 4-19 C) revealed a more

than 6 x higher Ki of the mutant, i.e. increased NFZ resistance relative to wild

type OsPDS-His6 (Table 6), as expected from reports on this mutation in

orthologous enzymes in vivo (see above). The DPQ concentration

dependency (Fig. 4-19 B) showed that the apparent KM for DPQ was strongly

decreased by factor of 3.5 (Table 6) indicating increased DPQ affinity.

Fig. 4-19 Kinetic characterization of Arg300Ser PDS.

(A, B) Concentration dependency of Arg300Ser PDS for phytoene (A) and DPQ (B).

(C) Inhibition of Arg300Ser PDS in a matrix of varying DPQ and NFZ concentrations.

Data points (squares, ζ-carotene; circles, phytofluene) represent the mean of

duplicates ± SEM. The deduced ζ-carotene/phytofluene ratio (dotted line; plotted to

the right Y axis) in A and B and is given as triangles. Data in A (R2 = 0.58) and B (R2 =

0.95) were fitted with the Michaelis-Menten equation and the equation for

Phytoene [mM]

v [n

mol

min

-1 m

g-1] z/pf [nm

ol/nmol]

0 10 20 30 400

1

2

3

4

5

6

0.0

0.2

0.4

0.6

DPQ [mM]

v [n

mol

min

-1 m

g-1] z/pf [nm

ol/nmol]

0.0 0.5 1.0 1.5 2.0 2.50.0

0.5

1.0

1.5

0.0

0.2

0.4

0.6

0.8

DPQ [mM]

v [n

mol

ζ m

in-1

mg-1

]

0.0 0.2 0.4 0.6 0.8 1.0 1.20.0

0.2

0.4

0.6

0.8

1.0

0.00 mM NFZ0.55 mM NFZ1.10 mM NFZ2.20 mM NFZ

A B

C

Page 100: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

86

competitive inhibition in C (R2 = 0.94). The ζ-carotene/phytofluene ratios were fitted

with a spline. Assays were incubated for 15 min under standard conditions.

Moreover, the phytoene concentration dependency (Fig. 4-19 A) supported

that KM for phytoene was lowered by a factor of 12 due to the mutation

(Table 6). Concomitantly, a strong decrease in Vmax was observed for both

phytoene and DPQ (Table 6), corroborating the loss of activity of the mutant.

The lowered Ki for NFZ accompanied by a decreased KM for DPQ is not fully

compatible with an equivalent binding mode of the two (see 5.1.6.2).

Table 6 Substrate and inhibitor affinities for wild type OsPDS-His6 and Arg300Ser PDS.

The Michaelis-Menten parameter values deduced from substrate-concentration

dependencies for the substrates 15-cis-phytoene and DPQ are summarized for both

PDS versions. * denotes that 10 mM phytoene instead of 40 mM phytoene were used

during DPQ concentration dependencies. Ki values were determined in quinone-

competitive inhibition kinetics.

Ligand KM or Ki

WT

[mM]

KM or Ki

Arg300Ser

[mM]

Vmax

WT

[nmol min-1 mg-1]

Vmax

Arg300Ser

[nmol min-1 mg-1]

DPQ 1.3 ± 0.2 0.4 ± 0.1 28.1 ± 1.4 1.2 ± 0.1*

Phytoene 53.9 ± 18.1 4.5 ± 2.6 46.3 ± 10.7 1.0 ± 0.2

NFZ 0.2 ± 0.0 1.5 ± 0.4 - -

A main difference in the phytoene and DPQ concentration dependencies is

that phytofluene release was predominant for Arg300Ser PDS (dotted lines in

Fig. 4-19 A, B); this is the opposite with the wild type enzyme (see 4.2.3).

Nevertheless, a similar trend can be noted with both versions: low phytoene

concentrations and high DPQ concentrations favor formation of the end

product ζ-carotene, with the opposite favoring release of the intermediate.

Page 101: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

87

In summary, the affinities for all ligands are strongly impaired in Arg300Ser PDS,

supporting multiple effects on the active site and major changes in enzyme

kinetics including the observed loss in enzymatic activity. In an attempt to

pinpoint sub-processes impaired in the mutant (such as altered DPQ

reoxidation), mathematical modelling was applied. However, this did not

allow an unequivocal identification.

Apart from Arg300Ser, other mutations that were reported to confer NFZ

resistance in orthologous PDS enzymes were introduced into OsPDS-His6 by

site-directed mutagenesis – namely Leu538Phe, Leu538Arg and Phe162Val. The

mutant enzymes were all purified for biochemical and kinetic

characterization. Like Arg300Ser, they all exhibited wild type-like flavinylation

and most likely a native overall fold but less than 5 % of the wild type activity.

This residual activity did not allow kinetic characterization.

4.3.3 Identification of non-covalent norflurazon – OsPDS-His6 interactions

Structure-activity correlations for NFZ derivatives were carried out in the 1980s

and 1990s in complex systems. This suggested that lipophilicity and electron-

withdrawing properties of the meta-trifluoromethyl substituent of the phenyl

moiety are crucial for the effectiveness of norflurazon (see 1.3.1.3 and

references therein). With the crystal structure of OsPDS-His6 in complex with

NFZ resolved, non-covalent interactions of the m-CF3 group with residues

could now be deduced from changes in electron density gradients by

quantum mechanical calculations (kindly provided by John Delaney,

Syngenta, Bracknell, UK) in order to characterize interactions that are unique

to this substituent and explain its inhibitory effectiveness.

The m-CF3-phenyl and its contact points and interactions in OsPDS-His6 are

visualized in Fig. 4-20 A and B. m-CF3-phenyl exhibits a large number of weakly

attractive non-covalent interactions rather than relying on a single dominant

one (Fig. 4-20 C). Some interactions occur for any phenyl ring, irrespective of

its substituents, e.g. with the carbonyl of FAD, the sulphur of Met310 or the

Page 102: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

88

phenyl ring of Phe162. Weakly attractive interactions that are unique for the

CF3 substituent and could only partially occur for other substituents occur

between the fluorine atoms with Met188, Met277, Ala280 and Phe423. Thus, the

CF3 substituent indeed strongly contributes to the affinity and potency of NFZ

by five weakly attractive interactions with the PQ binding site.

Fig. 4-20 Non-covalent interactions between m-CF3-phenyl in NFZ and OsPDS-His6.

Weakly attractive non-covalent interactions between the m-CF3-phenyl moiety of

NFZ and residues in OsPDS-His6 are given as dashed lines. Residue numbering refers to

O. sativa PDS (Acc. A2XDA1.2) including its N-terminal 87 amino acid transit peptide.

Protein alignments reveal that all of the above identified residues are strictly

conserved (Fig. 4-16), suggesting that the described interactions with m-CF3-

phenyl generally apply to PDS and CrtP enzymes in cyanobacteria, green

algae and plants. Amongst these residues, the mutation corresponding to

Phe162Val was reported to confer resistance to NFZ in Chlamydomonas

reinhardtii (Suarez et al., 2014). This supports the notion that these residues are

essential for the interaction between the target enzyme and its herbicidal

inhibitor and may represent targets in the development of optimized

herbicides and of herbicide resistant PDS enzymes.

CH

HCHH

S

O

H

H

F

F

FH

H

NH

H

S

NO

Cl

NH

F162

M188

F423

A280

M277

M310

FAD

Page 103: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

89

4.4 ζ-Carotene desaturase ZDS

4.4.1 Basic biochemical characterization of OsZDS-His6

Having investigated the structure and kinetics of purified PDS, we extended

our investigation to ZDS. PDS and ZDS are paralogs that have evolved through

an ancient gene duplication event. The similarities and differences of these

two desaturases regarding structure and enzyme kinetics are crucial for the

understanding of the poly-cis pathway of carotene desaturation.

4.4.1.1 Establishing native OsZDS-His6 purification and an in vitro assay

Heterologous expression in E. coli and native purification of ZDS-His6 from

Oryza sativa was established. Initially, cloning and purification for OsZDS-His6

was carried out using the procedures for OsPDS-His6 by Gemmecker et al.

(2015) as a guideline. However, the OsZDS-His6 yield per gram E. coli was as

low as 25 µg compared to 750 with OsPDS-His6. Extensive optimization of the

expression conditions, including the use of various E. coli strains, did not result

in much improvement. In the end, codon optimization and expression in the E.

coli strain Rosetta, harboring plasmid-encoded rare tRNAs, allowed obtaining

200 µg OsZDS-His6 g-1. During optimization of the purification, the addition of

0.01 % (w/v) TritonX100 during cell disruption and IMAC resin binding as well as

the addition of 10 mM MgCl2 throughout purification proved to improve ZDS

yields. The optimized procedures allowed obtaining 350 µg OsZDS-His6 g-1.

SDS-PAGE analysis of purified OsZDS-His6 (Fig. 4-21 A) revealed a dominant

protein band with a size of approximately 60 kDa (calculated mass of 60.3

kDa), accompanied by only minor impurities. The IMAC eluates appeared

yellow, suggesting the presence of a flavin. No flavin fluorescence was

detected upon SDS-PAGE indicating non-covalent association. In order to

prove its native state, enzymatic activity was assayed using the biphasic

liposomes-based system established for PDS (Gemmecker et al., 2015). The

hydrophobic ZDS substrate 9,9’-di-cis-ζ-carotene was incorporated into

liposomes and decylplastoquinone, the assumed exchangeable redox

cofactor of ZDS, was allowed to partition into liposomes.

Page 104: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

90

Fig. 4-21 SDS-PAGE analysis and in vitro activity assay for purified OsZDS-His6.

(A) SDS-PAGE analysis of purified OsZDS-His6. 25 µg of protein were separated on a 12

% polyacrylamide gel. PageRulerTM Prestained Protein Ladder (ThermoFisher) was

used as size marker. (B) OsZDS-His6 in vitro assays were analyzed by HPLC after

extraction. The UV/VIS spectra of the substrate (9,9’-di-cis-ζ-carotene) and the

formed products (proneurosporene, prolycopene) are given as insets. No products

were detected in the assay with the empty vector control (not shown).

HPLC analysis revealed the formation of the canonical ZDS products

proneurosporene (PN; 7,9,9’-tri-cis-neurosporene) and prolycopene (PL;

7,9,7’,9’-tetra-cis-lycopene). They can unequivocally be identified as the sole

ZDS products due to their unique UV/VIS spectra (Fig. 4-21 B). In accordance

with the procedures for PDS, ZDS remained active and soluble upon storage

at - 20 °C, provided that imidazole was removed by dialysis and replaced by

glycerol. Like with PDS, the presence of detergents after purification showed

to be detrimental to ZDS activity. They were omitted in all buffers subsequent

to enzyme binding to the IMAC resin. Thus, ZDS is similar to PDS as it behaves

as soluble during purification, remaining native in the absence of detergents,

and is capable of spontaneously attaching to liposomal membranes. Thus,

OsZDS-His6 represents a monotopic membrane protein.

A B

400 425 nm

379

1 2 4 5 6 7 8

min

mA

U(4

00

-45

0n

m) 409

432 nm

438 nm

prolycopene

proneurosporene9,9‘-di-cis-ζ-carotene

9

kDa

100

70

55

40

35

25

15

10

Page 105: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

91

4.4.1.2 Identification of nucleotide redox cofactors bound to OsZDS-His6

The conserved Rossmann fold of ZDS (Breitenbach et al., 1998) could enable

the binding of FAD and FMN as well as of other mono- or dinucleotides. To

detect these redox cofactors, purified OsZDS-His6 was heat-denatured to

release non-covalently bound enzyme cofactors. The supernatant obtained

upon centrifugation was subjected to LC-MS/MS analysis. Authentic standards

of FAD, FMN, NADH and NADPH were used as references (Fig. 4-22).

Fig. 4-22 Identification of mono- and dinucleotide cofactors bound to OsZDS-His6.

The supernatant of heat-denatured OsZDS-His6 was analyzed by LC-MS/MS and

nucleotide cofactors were identified using FAD, FMN, NADH and NADPH as authentic

standards. (A) Maxplot (280 – 800 nm) of ZDS supernatant and of an FAD – FMN

standard. UV/VIS spectra are given. (B) Extracted ion chromatogram (XIC) for the

A

0

20

40

60

80

100

Rel

ativ

e A

bun

dan

ce 11.41m/z = 457.10275

-457.12103

0

20

40

60

80

100

Rel

ativ

e A

bun

dan

ce 10.57m/z = 786.14869

-786.18013

10.57

11.42

300 350 400 450m/z

Rel

ativ

e A

bund

ance

439.10080348.06989

20

40

60

80

100

300 350 400 450m/z

Rel

ativ

e A

bund

ance

359.13440

396.09491

300 350 400 450m/z

Rel

ativ

e A

bund

ance

359.13449

396.09491

10.49

11.34

B

C

9 10 11 12 13min

mA

U(2

80-8

00n

m)

10.49

11.33

445 nm

FAD standard

FMN standard

FAD + FMN

447 nm447 nm

FMNFAD

ZDS

FAD

FMN

ZDS

FMN

ZDS

FAD

9 10 11 12 13

9 10 11 12 13

min

min

300 350 400 450m/z

Rel

ativ

e A

bund

ance

439.10080348.06989

20

40

60

80

100

20

40

60

80

100

20

40

60

80

100

FAD in ZDS

FMN in ZDS

Page 106: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

92

FAD [M+H]+ with a calculated m/z of 786.16441 ± 20 ppm for the OsZDS-His6

supernatant and the FAD standard. The corresponding MS2 spectra are given. (C)

Extracted ion chromatogram (XIC) for the FMN [M+H]+ with a calculated m/z of

457.11689 ± 20 ppm, for the OsZDS-His6 supernatant and the FMN standard. The

corresponding MS2 spectra are given.

As shown in Fig. 4-22 A, OsZDS-His6 mainly contained FAD and only trace

amounts of FMN. Centrifugation of the samples yielded a white ZDS pellet,

confirming non-covalent binding. As both flavins contain an isoalloxazine as

chromophore, the FAD/FMN ratio can be derived from peak areas at 447 nm

and was found to be 96 % FAD and 4 % FMN. The identity of these two

cofactors in OsZDS-His6 is supported by the presence of the expected

quasimolecular ions and by MS2 spectral characteristics in comparison to

authentic standards (Fig. 4-22 B and Fig. 4-22 C). No signals corresponding to

the [M+H]+ of NAD+, NADH, NADP+ or NADPH were detected. In conclusion,

OsZDS-His6 contains FAD as a non-covalently bound redox cofactor that can

occasionally be replaced by FMN.

In order to determine the percentage of flavinylation in OsZDS-His6

preparations, FAD and protein were photometrically quantified. Assuming

flavin binding to ZDS in a 1:1 stochiometry, 35 % of ZDS represented the

holoenzyme whereas 65 % were apoenzyme. Like with PDS (Gemmecker et

al., 2015), the addition of 25 µM FAD to assays did not improve OsZDS-His6

activity. Accordingly, ZDS cannot exchange FAD, at least in vitro. FAD can

therefore be assumed to be permanently bound to carry out multiple

reaction cycles (provided it is reoxidized by PQ) while the apoenzyme is most

likely irreversibly inactivated.

4.4.1.3 Kinetic mechanism of the OsZDS-His6 bi-substrate reaction

As laid out for PDS (4.1.4), the question whether quinone is required for

carrying out the carotene desaturation reaction in ZDS per se is of

mechanistic interest. OsZDS-His6 assays were carried out in the absence of

DPQ. In fact, product formation was observed in the absence of quinones:

Page 107: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

93

0.23 nmol of flavinylated ZDS monomers yielded 0.05 nmol PN and 0.04 nmol

PL, corresponding to 0.13 nmol of double bonds formed which is close to

equimolarity. Thus, like with PDS, carotene desaturation and PQ reduction are

separate catalytic events. Each flavinylated OsZDS-His6 monomer can be

assumed to perform one catalytic cycle that is independent of DPQ. The

latter is solely needed kinetically to regenerate oxidized FAD for repeated

catalytic cycles. This supports the prevalence of an ordered ping pong bi bi

mechanism, like with OsPDS-His6 (see 4.1.4).

4.4.2 Structure and reaction mechanism of OsZDS-His6

4.4.2.1 Reaction mechanism of carotene desaturation by OsZDS-His6

As shown in this work, OsPDS-His6 does not employ an acid-base component

in its reaction mechanism for carotene desaturation, as witnessed inter alia by

the lack of deuterium incorporation from heavy water into its products.

Rather, a “flavin only” reaction mechanism based on the redox-active

isoalloxazine of FAD as sole catalyst prevails (for details, see 4.1.3).

In order to test the hypothesis that ZDS does not rely on acid-base catalysis

either, deuterium incorporation into PL formed by OsZDS-His6 in the presence

of D2O was monitored by LC-MS (Fig. 4-23). PL was formed from 9,9’-di-cis-ζ-

carotene in buffers containing either D2O or H2O, as witnessed by the same

UV/VIS spectrum (Fig. 4-23 A), the same [M+H]+ mass of 537.44548 ± 5 ppm

corresponding to C40H57 (Fig. 4-23 B) and the same mass for the M+1 signal that

are due to 13C incorporation (Fig. 4-23 C). Consequently, there is no

deuterium incorporation from an outer source into products formed by OsZDS-

His6. This does not imply acid-base catalysis but supports a “flavin only”

reaction mechanism, as postulated for PDS (see 4.1.3).

Page 108: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

94

Fig. 4-23 LC-MS analysis of ζ-carotene formed in buffer containing heavy water.

(A) The formation of PL was detected at 350 - 500 nm upon incubation in reaction

buffers with D2O and H2O. The UV/VIS spectrum of the PL formed is given as inset. (B)

Extracted ion chromatogram (XIC) for [M+H]+ of PL with a calculated m/z ratio of

537.44548 ± 5 ppm, verifying its formation in the absence and the presence of D2O.

(C) The corresponding MS1 spectra of PL formed with H2O (left) and with D2O (right) in

reaction buffers reveal identical [M+H]+ and M+1 signals.

4.4.2.2 Homology modeling of OsZDS-His6 and analysis of the active site

To further investigate structural and functional differences between PDS and

ZDS, X-ray protein crystallography for OsZDS-His6 was anticipated. However, all

attempts during this thesis failed despite much variation of the protocols. We

therefore resorted to homology modeling using its paralog OsPDS-His6 (33 %

sequence identity) as modeling template. In silico docking of riboflavin into

OsZDS-His6 was performed in order to define the active center.

According to this, the structure of the paralogous carotene desaturases

OsPDS and OsZDS is strongly conserved, with ZDS also containing a conserved

Rossman fold, a HotDog domain and a membrane binding domain (Fig. 4-24

A). One long, hydrophobic substrate cavity originates from the membrane

binding domain surface (Fig. 4-24 A), providing access to the FAD-containing

436 nm

11 12 13min

12.3812.39

D2O

H2O

C40H57 (prolycopene)

m/z = 537.44441-541.44655

Re

l. ab

und

ance

A B

C

11 12 13min

12.34

D2OH2O

mA

U(3

50-

500

nm) 13.05

prolycopene

proneurosporene

536 538 540 542 544m/z

0102030405060708090

10012.29-12.45 min

ZDS + H2O537.44589

538.44924

[M+H

]+

M+

1

536 538 540 542 544

m/z

0102030405060708090

100537.44588

538.44958

12.29-12.45 minZDS + D2O

[M+H

]+

M+

1

Page 109: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

95

active center for the lipophilic, membrane-soluble carotene substrates and

PQ. Like with OsPDS-His6, the existence of only one substrate cavity supports

an ordered ping pong bi bi mechanism. This is corroborated by the fact that

carotene desaturation by OsZDS-His6 occurred in the absence of PQ (see

4.4.1.3). As expected, the isoalloxazine is predicted to reside within the

substrate cavity whereas the D-ribitol moiety is accommodated in a cavity of

the FAD-binding domain (Fig. 4-24 A) with the Gly-Ala-Gly (GAG) motif of the

Rossman fold coordinating the phosphates in FAD (Hanukoglu, 2015). Like in

PDS, the isoalloxazine position in fact is likely to be shifted into the side cavity

(Fig. 4-24 A and B). The isoalloxazine environment defines the active center.

Accordingly, the redox-active benzoquinone moiety of DPQ was docked into

the active site of OsZDS-His6 apoenzyme (not shown). Interestingly, the

inhomologous bacterial phytoene desaturase CrtI possesses a similar tertiary

structure (Fig. 4-24 A).

Fig. 4-24 Prediction of the OsZDS structure by homology modeling.

(A) Structural comparison of monomeric Oryza sativa PDS, Oryza sativa ZDS and CrtI

from Pantoea ananatis. The structure of OsZDS was predicted by homology

Asp149

Arg148

Arg152

Asp489

Glu492

His115

His86

Glu451

His159

Arg300

Glu542

Lys142

OsPDS OsZDSA

B

FAD binding

domainMembrane

binding domain

HotDog domainCrtI

OsPDS CrtIOsZDS

Page 110: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

96

modeling using the OsPDS structure as modeling template. The riboflavin position

results from in silico docking. The structures are depicted as cartoon, with FAD (PDS),

riboflavin (ZDS) and isoalloxazine (CrtI) represented by orange sticks and the inner

surface of the hydrophobic substrate cavity given in grey. The assumed entry to the

substrate cavity in the membrane binding domain is indicated by an arrow. (B)

Active center of the OsPDS, OsZDS and CrtI. All basic and acidic amino acid residues

within a 8 Å radius around the redox-reactive isoalloxazine (orange sticks) are given

as sticks, color coded by elements (blue, nitrogen; red, oxygen; grey, carbon). The

GAG motif of the Rossmann fold is given as dark grey sticks (right bottom).

Numbering of residues in OsZDS refers to the mature OsZDS protein (Acc.

XP_015646524.1) including the addition of a N-terminal methionine. Numbering in

OsPDS relates to the immature protein including its N-terminal 87 amino acid transit

peptide (Acc. A2XDA1.2). Numbering in CrtI is as given in Schaub et al. (2012).

The OsZDS active center environment (defined as 8 Å radius around the

isoalloxazine) shows only few basic or acidic amino acids – namely His86, His115,

Glu492 and Asp489 (Fig. 4-24 B) which are conserved in cyanobacterial CrtQb

and plant ZDS enzymes (not shown) and might initiate carotene desaturation

by acid-base catalysis. However, as with OsPDS-His6 (see 4.1.3), the lack of

deuterium incorporation during carotene desaturation by OsZDS-His6 (Fig. 4-

23) suggests that no acid-base catalysis is employed. This is in accordance

with the acidic and basic residues being too distant from the carotene

substrate cavity for direct interaction with the carotene reaction site (Fig. 4-24

B). To further illustrate this, the active center of CrtI, employing acid-base

catalysis (Schaub et al., 2012) is depicted (Fig. 4-24 B). It drastically differs from

OsPDS and OsZDS as it contains two arginines and one aspartate as catalytic

triad in less than 5 Å distance of the isoalloxazine and the carotene binding

site (Schaub et al., 2012). Conclusively, these findings imply that ZDS, like PDS,

employs a “flavin only” reaction mechanism with FAD acting as sole catalyst

(see 4.1.3; Brausemann et al., 2017).

Page 111: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

97

4.4.2.3 Homooligomerization of OsZDS-His6 in solution

OsPDS-His6 is only active when assembled as a homotetramer containing one

FAD per subunit whereas monomers are unflavinylated and inactive

(Gemmecker et al., 2015, see 4.1.1.2). This prompted us to revisit the results of

Albrecht et al. (1995) who suggested ZDS from Capsicum anuum to be active

in monomeric form.

Fig. 4-25: Analysis of purified OsZDS-His6 on size standard-calibrated GPC columns.

(A, B) GPC profile of OsZDS-His6 on a Superose 6 10/300 GL column, when purified

and analyzed in the absence (A) or in the presence of TCEP (B). Protein absorption

was monitored at 280 nm, FAD fluorescence at 530 nm upon excitation at 450 nm.

Two OsZDS-His6 populations of different molecular mass were detected (I and II). (C)

The partition coefficient KAV was determined for size standard proteins and OsZDS-His6

samples (the latter given in red) on three GPC columns differing in fractionation

ranges and plotted against the logarithm of the molecular weight logMW [kDa]. Size

marker proteins ranged from 12.4 kDa to 440 kDa. Analysis of the OsZDS-His6

indicated that I corresponds to a monomer (60.3 kD) and II to a dimer (120.6 kDa).

Purified OsZDS-His6 was subjected to GPC analysis. In accordance with the

results obtained with PDS, imidazole was essential to prevent OsZDS-His6 from

aggregation and adsorption to the column resin. GPC analysis on a Superose

6 10/300 GL column revealed two major populations when TCEP was omitted

10 100 1000

0.2

0.4

0.6

0.8HiLoad 16/60

Superdex 75

Superose 6

0 mM TCEP

1 mM TCEP

Em530nm

A280nm

A280nm

A

B

5 10 15 20

5 10 15 200 ml

ml

VE

VE

C

logMW [kDa]

KA

V

I

I: 60.3 kDa

II: 120.6 kDa

II I

I

free FAD

mA

Um

AU

0

Page 112: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

98

during purification and GPC (I and II in Fig. 4-25 A). Additionally, the higher

mass population II showed a leading peak indicating higher order OsZDS-His6

associates. When TCEP was present throughout purification and GPC or when

OsZDS-His6 was incubated with 1 mM TCEP for 5 min after purification, only the

lower mass population I was observed (Fig. 4-25 B). The determination of the

partition coefficient KAV on calibrated columns revealed that the low mass

population I corresponded to the monomer (60.3 kDa) while the high mass

population II corresponded to the dimer (120.6 kDa) of OsZDS-His6 (Fig. 4-25

C). Dimerization and higher order oligomerization under oxidizing conditions

suggested the participation of intermolecular disulfide bonds. However, SDS-

PAGE under oxidizing conditions yielded only the monomer band of 60 kDa

(not shown), indicating oligomerization to rely on non-covalent interactions.

GPC analysis revealed that the OsZDS-His6 monomer contained FAD as

witnessed by flavin fluorescence (Fig. 4-25 B, bottom panel), suggesting that

OsZDS-His6 might in fact be active as monomer. Two experimental

approaches were chosen to further investigate the dependence of ZDS

activity on homooligomerization. First, OsZDS-His6 was purified in the presence

and absence of TCEP, yielding a purely enzyme samples of different

oligomerization (Fig. 4-25 A). Both showed identical flavinylation percentage

(35 %) and similar specific activity (0 mM TCEP: 311 pmol PL and 147 pmol PN

min-1 mg-1; 1mM TCEP: 275 pmol PL and 166 pmol PN min-1 mg-1). Second, the

specific activity of the isolated dimeric and monomeric GPC fractions (Fig. 4-

25 A) were determined individually, revealing that the dimer was 5 x more

active than the monomer (monomer: 19 pmol PL and 10 pmol PN min-1 mg-1;

dimer: 94 pmol PL and 33 pmol PN min-1 mg-1).

In conclusion, OsZDS-His6 can be enzymatically active as a monomer with the

dimer potentially fostering activity. In contrast, the homotetramer represents

the minimal oligomer required for catalytic activity in OsPDS-His6, while

monomers are inactive (Gemmecker et al., 2015).

Page 113: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

99

4.4.3 Kinetic characterization of OsZDS-His6

4.4.3.1 Basic characterization and optimization of reaction parameters

The OsZDS-His6 assay conditions were optimized regarding the added amount

of enzyme, temperature and pH (Fig. 4-26). As a starting point, assays

contained 5 mM 9,9’-di-cis-ζ-carotene and 19.25 mM DPQ and were

incubated for 30 min.

Fig. 4-26 Basic characterization of the OsZDS-His6 reaction.

(A) protein, (B) pH and (C) temperature dependency of the OsZDS-His6 reaction.

Circles, PL; squares, proneurosporene. Each experiment was carried out using the

optimum value of the respective non-variable parameter – i.e. 36 °C and pH 6.5 (A),

pH 6.5 and 25 µg ZDS (B) and 36°C and 25 µg ZDS (C). Incubation time was 30 min.

Data in (B) and (C) represent the mean of triplicates ± SEM. Measurements in (A)

were carried out once. Data were fitted with a spline.

Protein linearity for PL formation was observed up to ca. 100 µg OsZDS-His6 per

assay and up to 50 µg for the intermediate PN (Fig. 4-26 A). Thus, as observed

with OsPDS-His6, increasing protein concentrations favored end product

ZDS [µg]

v [p

mol

min

-1]

0 25 50 75 1000

5

10

15

temperature [°C]

v [p

mol

min

-1 m

g-1

]

24 28 32 36 400

200

400

600

800

pH

v [p

mol

min

-1 m

g-1

]

3.5 4.0 4.5 5.0 5.5 6.0 6.5 7.0 7.50

250

500

750

1000

A B

C

Page 114: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

100

formation. Regarding temperature, PN and PL formation responded similarly

with the reaction velocity increasing up to 40 °C (Fig. 4-26 B). PN formation

showed a broad pH optimum around pH 5.5, whereas the optimum of PL

formation was shifted towards pH 6.5 – pH 7.0. Thus, the velocity of PL end

product formation dominated at pH values above 6.0 (Fig. 4-26 C). These

data define the standard conditions for OsZDS-His6 assays: 25 µg ZDS per assay

(35 µg ml-1), pH 6.5, 36 °C, 5 mM ζ-carotene and 19.25 mM DPQ (the latter in

analogy to PDS). The incubation time was 30 min.

4.4.3.2 Reaction time course of ζ-carotene conversion by OsZDS-His6

The conversion of 9,9’-di-cis-ζ-carotene into the products PN and PL was

monitored by HPLC (Fig. 4-27). The disappearance of the substrate

proceeded over 90 min (Fig. 4-27 B) and ceased after 360 min. Notably, 9-cis-

ζ-carotene was present in liposomes due to non-enzymatic isomerization and

was barely converted (Fig. 4-27 B). The formation of the intermediate PN

preceded the formation of the end product PL during the first 5 minutes (Fig.

4-27 A). Towards longer incubation than 90 min, the intermediate

approached a steady state situation while PL accumulated steadily. As for

PDS (see 4.2.1), this is typical of two-step reactions catalyzed by two enzymes

or bifunctional enzymes like OsZDS-His6 (McClure et al., 1969).

Fig. 4-27 Reaction time course of proneurosporene and prolycopene formation from

ζ–carotene by OsZDS-His6.

(A) Reaction time course of product formation and (B) reaction time course of

substrate decrease. The OsZDS-His6 reaction was conducted under standard

time [min]

caro

tene

[mM

]

0 20 40 60 800.0

0.1

0.2

0.3

0.4

time [min]

caro

tene

[mM

]

0 20 40 60 801

2

3

4

5A B

Page 115: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

101

conditions. Filled circles, PL; filled squares, proneurosporene; open circles, 9-cis-ζ-

carotene; open squares, 9,9’-di-cis-ζ-carotene. Data represent the mean of

triplicates ± SEM and were fitted with a spline.

Like with OsPDS-His6 (see 4.2.1), OsZDS-His6 activity ceased after 360 min due

to enzyme inactivation. 83 % end product formation (PL and cis-lycopenes

originating from unspecific isomerization) could be attained upon repeated

addition of fresh OsZDS-His6, and long-time incubation, with only 11 % of PN

and 6 % of the substrate 9,9’-di-cis-ζ-carotene not being converted., Thus, the

equilibrium of ζ-carotene desaturation by OsZDS-His6 is far on the product

side, as with OsPDS-His6 (see 4.2.1)

4.4.3.3 Substrate concentration dependency of the OsZDS-His6 reaction

The concentration dependency on 9,9’-di-cis-ζ-carotene and

proneurosporene was examined under DPQ saturation (19.25 mM ≈ 100 KM;

Fig. 4-28). With 9,9’-di-cis-ζ-carotene as substrate, the characteristics of

intermediate and end product formation differed substantially (Fig. 4-28 A).

Fitting with the Michaelis-Menten equation for PL formation revealed an

apparent KM of 5.6 ± 0.7 mM 9,9’-di-cis-ζ-carotene and a Vmax of 609 ± 23

pmol min-1 mg-1 PL. Saturation was reached at low ζ-carotene

concentrations. In contrast, the PN formation steadily increased so that KM

and Vmax could not be determined. This inequivalent kinetic behavior of PN

and PL formation favored intermediate release at high substrate

concentrations, i.e. the PL/PN ratio decreased from 1.5:1 to 0.4:1 with

increasing concentrations of ζ-carotene (Fig. 4-28 A). This is in accordance

with the findings with OsPDS-His6 (see 4.2.3). With PN as substrate, fitting the

formation of PL with the Michaelis-Menten equation yielded an apparent KM

of 14 ± 3 mM PN and a Vmax of 6319 ± 662 pmol min-1 mg-1 PL (Fig. 4-28 B).

Lastly, the dependency of the OsZDS-His6 reaction rate on DPQ concentration

was examined under ζ-carotene saturation (10 mM ≈ 2 KM). Again, there was

kinetic inequivalence in the formation of PL and PN (Fig. 4-28 C). PL formation

can be fitted with the Michaelis-Menten equation revealing a KM of 0.3 ± 0.1

Page 116: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

102

mM and a Vmax of 813 ± 32 pmol min-1 mg-1 PL whereas the formation of the

intermediate PN, being in steady state, barely follows Michaelis-Menten

kinetics. While PN formation dominates at low DPQ concentrations, high DPQ

supply favors end product formation with a maximum PL/PN ratio of 1:1.

Moreover, this supports that the standard DPQ concentration used (19.25

mM) is by far saturating.

In summary, high DPQ concentrations and low ζ-carotene fluxes through ZDS,

i.e. low ζ-carotene concentrations, are needed to favor the formation of the

end product PL over the intermediate PN. This is in accordance with the

findings with OsPDS-His6 (see 4.2.3).

Fig. 4-28 Substrate concentration dependencies of the OsZDS-His6 reaction rates.

(A – B) Dependency of OsZDS-His6 on the concentration of ζ-carotene (A) and

proneurosporene (B), conducted at a fixed DPQ concentration (19.25 mM; ≈ 100 x

KM). (C) Dependency of OsZDS-His6 on the DPQ concentration, conducted at a fixed

ζ-carotene concentration (10 mM; ≈ 2 x KM). Circles, PL; squares, proneurosporene;

triangles, PL/PN ratio. Data represent the mean of triplicates ± SEM. The PL:PN ratios in

9,9'-di- cis-ζζζζ-carotene [mM]

v [p

mol

min

-1 m

g-1

]

PL / P

N ratio [m

ol/mol]

0 5 10 15 20 250

400

800

1200

0

2

4

6

proneurosporene [mM]

v [p

mol

min

-1 m

g-1

]

0.0 2.5 5.0 7.5 10.0 12.5 15.00

1000

2000

3000A B

DPQ [mM]

v [p

mol

min

-1 m

g-1

]P

L / PN

ratio [mol/m

ol]

0.0 0.5 1.0 1.5 2.0 2.50

250

500

750

0.0

0.5

1.0

1.5C

Page 117: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

103

A and C (dotted lines) are plotted to the right y-axis. The data were fitted with the

Michaelis-Menten equation (goodness of fit for PL formation: A, R2 = 0.99; B, R2 = 0.99;

C, R2 = 0.97). PL/PN ratios were fitted with a spline.

4.5 ζ-Carotene isomerase ZISO

4.5.1 Substrate specificity of OsZISO-His6 in E. coli

As laid out in 1.3.2, ZISO is the most recently identified and least examined

enzyme involved in the poly-cis pathway of carotene desaturation. It is the

only integral membrane protein in the pathway and shares no homology with

other plant enzymes (Chen et al., 2010). ZISO catalyzes the isomerization of

the PDS product 9,15,9’-tri-cis-ζ-carotene to form the ZDS substrate 9,9’-di-cis-

ζ-carotene when co-expressed with PDS in E. coli (Li et al., 2007; Chen et al.,

2010). Interestingly, PDS is an isomerase itself. Concomitant with desaturation,

PDS isomerizes the C9-C10 double bond from trans to cis, yielding 9,15,9’-tri-

cis-ζ-carotene from 15-cis-phytoene (Fig. 4-29). Considering this, we reasoned

whether ZISO is really an independent, bona fide enzyme. Alternatively, ZISO

might represent a membrane-integral scaffolding protein that interacts with

PDS, thereby enabling PDS to isomerize the C15-C15’ double bond. Thus, the

in vivo assay of Chen et al. (2010) is not sufficient for a definite

characterization of ZISO as enzyme. In order to address this question, we

aimed at establishing expression of OsZISO-His6 in E. coli and an in vitro assay

for ζ-carotene isomerization in the presence of OsPDS and in its absence.

First attempts of functional ZISO expression in ζ-carotene producing E. coli

failed because no isomerization of tri-cis-ζ-carotene to di-cis-ζ-carotene was

observed (“ζ-carotene + control / ZISO” in Fig. 4-29 A). Since ZISO is a

predicted transmembrane protein, it might not have integrated into

membranes in the absence of an appropriate signal peptide. Consequently,

the Mistic protein from Bacillus subtilis, a membrane insertion facilitator

(Roosild et al., 2005), was fused to the N-terminus of OsZISO-His6 via a flexible

peptide linker. In fact, Mistic-OsZISO-His6 proved to be functional in E. coli and

Page 118: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

104

9,15,9’-tri-cis-ζ-carotene was almost completely isomerized to 9,9’-di-cis-ζ-

carotene (“ζ-carotene + Mistic-ZISO” in Fig. 4-29 A).

Fig. 4-29 Cis-to-trans isomerization of the 15-cis double bond in ζ-carotene,

phytofluene and phytoene in E. coli upon coexpression of PDS and ZISO.

(A) Isomerization of 9,15,9’-ζ-carotene to 9,9‘-di-cis-ζ-carotene and (B) of 9,15-di-cis-

phytofluene to 9-cis-phytofluene upon co-expression of PDS and Mistic-OsZISO-His6.

(C) Isomerization of 15-cis-phytoene to all-trans-phytoene upon expression of Mistic-

A

min

min

mA

U40

0nm

0 1 2 3 4 5 6 7 8 9

9,9‘-di- cis-ζ-carotene

9,15,9‘-tri- cis-ζ-carotene

ζ-carotene + Mistic-ZISO

ζ-carotene + control / ZISO

398 nm422

377

295

9,15,9‘-tri- cis-ζ-carotene

400 nm424

379

0 1 2 3 4 5 6 7 8 9

mA

U33

1nm

phytofluene + Mistic-ZISO

phytofluene + control / ZISO9,15,9‘-tri-cis-ζ-carotene

9,9‘-di-cis-ζ-carotene

348 nm366

331

9-cis-phytofluene

347 nm366

331

254

9,15-di- cis-phytoflueneB

285 nm

15-cis-phytone

0 1 2 3 4 5 6 7 8 9min

mA

U28

5nm

15-cis-phytoene /all- trans-phytoene

phytoene + Mistic-ZISO

phytoene + control / ZISO

285 nm

all- trans-phytoeneC

9,9‘-di- cis-ζ-carotene

Page 119: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

105

OsZISO-His6. The UV/VIS spectra of the substrates and products are given as insets.

“Mistic-ZISO” denotes the expression of Mistic-OsZISO-His6 in E. coli, “control / ZISO

“denotes the expression of the empty vector or of unfunctional OsZISO-His6 in E. coli.

9,15,9‘-tri-cis-ζ-carotene was obtained from PDS assays, 9,9‘-di-cis-ζ-carotene from

tangerine tomato fruit. The 15-cis-/all-trans-phytoene mixture was obtained by

exposing purified 15-cis-phytoene to light. HPLC system 1 was used for separation.

Interestingly, also 9,15-di-cis-phytofluene that accumulated as a precursor of

tri-cis-ζ-carotene, was isomerized to presumably 9-cis-phytofluene (Fig. 4-29 B).

In contrast, expressing ZISO in 15-cis-phytoene-producing E. coli did not result

in detectable isomerization activity.

4.5.2 Reconstitution of the poly-cis pathway of carotene desaturation

The functional expression of ZISO allows revisiting the results of Bartley et al.

(1999). The partial reconstitution of the poly-cis pathway of carotene

desaturation achieved by expressing AtPDS and AtZDS in phytoene-

producing E. coli revealed photoisomerization of ζ-carotene as a rate limiting

step existing between the PDS and ZDS reaction. At the time, the authors

hypothesized that efficient carotenogenesis required the existence of a ζ-

carotene isomerase – now probably identified as ZISO by Li et al. (2007). To

evaluate whether photoisomerization of the photolabile 15-cis double bond in

9,15,9`-ζ-carotene would be as effective as ZISO in driving the pathway flux

through PDS and ZDS towards completion, i.e. PL formation, ζ-carotene isomer

composition in E. coli was assessed in the presence of either ZISO, in the dark

or under continuous light during growth (Fig. 4-30). Analysis of the carotene

patterns revealed that OsPDS and OsZDS in the dark led to the predominant

accumulation of 9,15,9’-tri-cis-ζ-carotene and 9,15-di-cis-phytofluene (Fig. 4-

30 “control”). Only minor amounts of the ZDS substrate 9,9’-di-cis-ζ-carotene

and of PL were formed. Upon strong light exposure, the significant amounts of

9,9’-di-cis-ζ-carotene formed by photoisomerization allowed substantial

formation of PL by OsZDS. Furthermore, strong unspecific isomerization of 9,15-

di-cis-phytofluene to presumably 9-cis-phytofluene and of PL to several

Page 120: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

106

unspecific cis-lycopene isomers was observed. In the presence of Mistic-

OsZISO, only a small proportion of tri-cis-ζ-carotene was left in the dark and

9,9’-di-cis-ζ-carotene was converted to PL by OsZDS even more effectively

than upon photoisomerization (compare ratio of PL/9,9’-di-cis-ζ-carotene for

“light” and “Mistic-ZISO” in Fig. 4-30). Moreover, almost no unspecific

isomerization of PL and phytofluene was observed.

Fig. 4-30 Pathway flux from 15-cis-phytoene to prolycopene in E. coli upon co-

expression of OsPDS, OsZDS and Mistic-OsZISO and in dependency on light.

Control, OsPDS and OsZDS were co-expressed and E. coli was grown in the dark;

Light, OsPDS and OsZDS were co-expressed and E. coli was grown under continuous

white light (800 µmol photons m-2 sec-1); Mistic-OsZISO, OsPDS, OsZDS and Mistic-

OsZISO were coexpressed and E. coli was grown in the dark. 1, lycopene isomers; 2,

prolycopene; 3, proneurosporene; 4, 9,15,9‘-tri-cis-ζ-carotene; 5, 9,9‘-di-cis-ζ-

carotene; 6, 9,15-di-cis-phytofluene; 7, 9-cis-phytofluene. Carotenes were identified

based on their UV/VIS spectra, given in Fig. 4-32 and Fig. 4-33. HPLC system 3 was

used for separation.

4.5.3 Isomerization of 9,15,9’-tri-cis-ζ-carotene by ZISO in vitro

As laid out in 4.5.1, ZISO can only be postulated as a bona fide enzyme if it

isomerizes ζ-carotene in vitro, in the absence of PDS and all other

carotenogenic enzymes. Therefore, having achieved functional ZISO

15P

control

Mistic-ZISO

light

5 6 7 8 9 10 11 12 13

mA

U30

0-50

0 nm

min

1 2 3 4 5 6 7

Page 121: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

107

expression in E. coli, a liposomal ZISO in vitro assay was established using

Mistic-ZISO-containing cell lysate as enzyme source.

First, Mistic-ZISO was added to PDS assays to show that Mistic-ZISO remains

active after cell disruption. Accordingly, as given in Fig. 4-31, the tri-cis-ζ-

carotene produced by PDS was readily and specifically isomerized to the di-

cis isomer (see “PDS assay” and PDS assay + Mistic-ZISO” in Fig. 4-31). When

omitting PDS and directly incorporating the ZISO substrate 9,15,9’-tri-cis-ζ-

carotene into liposomes, isomerization of the 15-cis double bond to trans

occurred as well – although to a smaller extent than in the presence of PDS

(Fig. 4-31). This observation unequivocally shows that ZISO is an isomerase.

Fig. 4-31 Cis-to-trans isomerization of the 15-cis double bond of 9,15,9’-tri-cis-ζ-

carotene by Mistic-OsZISO-His6 in vitro.

The isomerization of 9,15,9’-tri-cis-ζ-carotene was assessed by HPLC under the

following conditions: PDS, PDS in vitro assay yielding 9,15,9’-tri-cis-ζ-carotene from 15-

cis-phytoene; PDS assay + Mistic-ZISO, PDS in vitro assay upon the initial addition of

Mistic-OsZISO-His6-containing cell lysate;. 9,15,9’-tri-cis-ζ-carotene, incubation of

liposomes containing 9,15,9’-tri-cis-ζ-carotene; 9,15,9’-tri-cis-ζ-carotene + Mistic-ZISO,

incubation of liposomes containing 9,15,9’-tri-cis-ζ-carotene upon the addition of

Mistic-OsZISO-His6-containing cell lysate. The UV/VIS spectra of the substrate and

product are given as insets. HPLC system 1 was used for separation.

0 1 2 3 4 5 6 7 8 9

mA

U400nm

PDS assay

400 nm424

379

398 nm422

377

295

PDS assay + Mistic-ZISO

9,15,9‘-tri- cis-ζ-carotene

9,15,9‘-tri- cis-ζ-carotene + Mistic- ZISO

min

9,15,9‘-tri- cis-ζ-carotene9,9‘-di- cis-ζ-carotene

Page 122: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

108

At the time of obtaining these results, Beltrán et al. (2015) reported on the

native purification of ZISO, its independent enzymatic activity, its cofactor

requirements and postulated a reaction mechanism for heme b – dependent

isomerization. Given this situation, further research on ZISO in the framework of

this thesis was discontinued.

4.6 Apocarotenoids derived from the poly-cis pathway of carotene

desaturation

Acyclic apocarotenoids that might originate from carotenoid cleavage

dioxygenase (CCD)-mediated cleavage of poly-cis-configured carotene

desaturation intermediates are thought to be involved in the regulation of

carotenogenesis (Kachanovsky et al., 2012) and plant development

(Avendaño-Vázquez et al., 2014). If true, this might shed light on the

conundrum as to why plants have evolved a complex poly-cis pathway of

carotene desaturation (see 1.3). However, both of these studies are based on

observations made in plant mutants and lack the biochemical evidence for

the existence, let alone the identity of regulatory acyclic apocarotenoids.

Moreover, the cleavage specificity of CCDs for poly-cis-configured acyclic

carotenes has not been investigated.

Having established in vitro assays in our laboratory for all canonical CCDs from

Arabidopsis – namely AtCCD1, 4, 7 and 8 – we aimed at establishing the

purification of all plant-specific canonical cis-configured carotene

desaturation intermediates (see 1.3 and Fig. 1-2) and of some non-canonical

isomers in order to assay their CCD-mediated cleavage in vitro. This should

allow identifying yet unknown acylic apocarotenoids, representing potential

signaling molecules (or their precursors) whose existence has been postulated

(Kachanovsky et al., 2012; Avendaño-Vázquez et al., 2014). Carotene

desaturation intermediates are water-insoluble, plastid-localized C40

carotenes. Consequently, a CCD catalyzing the primary carotenoid

Page 123: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

109

cleavage en route to the postulated regulatory apocarotenoid is expected

to reside in plastids and to be capable of converting C40 carotenes. AtCCD4

and AtCCD7 fulfill both prerequisites (Bruno et al., 2016; Bruno et al., 2014) and

were therefore investigated as candidates for primary cleavage of acyclic

poly-cis carotenes (see 4.2.1 and 4.2.2). In contrast, AtCCD1 and AtCCD8 are

not candidates and were not investigated: AtCCD1 resides in the cytosol

(Walter and Strack, 2011) and AtCCD8 is highly specific for 9-cis-β-apo-10’-

carotenal, the substrate for strigolactone biosynthesis (Alder et al., 2012).

4.6.1 AtCCD4-mediated cleavage of acyclic poly-cis carotenes

Mutant analysis and phenotypic observations by Avendaño-Vázquez et al.

(2014) suggest that cis-ζ-carotene or cis-phytofluene is cleaved by AtCCD4 in

Arabidopsis. Thus, AtCCD4 cleavage activity was first tested using four ζ-

carotene isomers as substrates (Fig 1 A). These were the canonical PDS

product 9,15,9’-tri-cis-ζ-carotene and 9,9’-di-cis-ζ-carotene formed in planta

by ZISO. Furthermore, 9-cis-ζ-carotene was used since other 9-cis-configured

carotenoids give rise to signaling molecules such as ABA and strigolactones

(see 1.2.3) and all-trans-ζ-carotene was assayed as well. However, none of

these substrates was cleaved (Fig. 4-32 A), in contrast to the positive control

that relied on the well documented cleavage of β-apo-8-carotenal (Bruno et

al., 2016; not shown). Likewise, the four phytofluene isomers used were not

converted by AtCCD4: the canonical PDS intermediate 9,15-di-cis-

phytofluene, 9-cis-, 15-cis- and all-trans-phytofluene (Fig. 4-32 B).

Consequently, these in vitro data do not support the notion of Avendaño-

Vázquez et al. (2014) that a regulatory apocarotenoid originates from these

carotenes by CCD4-mediated primary cleavage.

Page 124: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

110

0 5 10 15 20 25 30 35 40

con

all-trans-ζ-carotene

AtCCD4

mA

U 4

00-4

50 n

m

5 10 15 20 25 30 35 40

con

9-cis-ζ-carotene

AtCCD4

5 10 15 20 25 30 35 40

con

9,9‘-di-cis-ζ-carotene

AtCCD4

A

con

9,15,9‘- tri-cis-ζ-carotene

5 10 15 20 25 30 35 40

AtCCD4

400 nm424

379

400 nm424

379

400 nm424

379

398 nm422

377

295286

0 5 10 15 20 25 30 35min

mA

U 3

50 n

m

con

9,15-di-cis-phytofluene

AtCCD4

con

AtCCD4

9-cis-phytofluene

conAtCCD4

all-trans-phytofluene

B 347 nm

366331

254

348 nm366

3315 10 15 20 25 30 35

5 10 15 20 25 30 35

min

Page 125: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

111

Fig. 4-32 HPLC analysis of AtCCD4 activity with carotene desaturation intermediates.

(A) Cleavage of (A) ζ-carotene isomers, (B) phytofluene isomers, (C)

proneurosporene, (D) prolycopene and (E) 9’-cis-neurosporene by AtCCD4. The

UV/VIS spectra of the substrates are given as insets. β-apo-8-carotenal served as a

positive control for AtCCD4 activity and was readily cleaved (not shown). HPLC

system 7 was used for A – D, HPLC system 9 was sued for E. For structures of substrates,

see Fig. 4-37.

Taking also into account that Kachanovsky et al. (2012) suggested a cis-

neurosporene- or cis-lycopene-derived signal, AtCCD4 activity with the

canonical ZDS products 7,9,9’-tri-cis-neurosporene (proneurosporene) and

7,9,7’,9’-tetra-cis-lycopene (prolycopene) was investigated. Moreover, 9’-cis-

neurosporene – produced as a byproduct from proneurosporene by CRTISO

(Isaacson et al., 2004; Yu et al., 2011) – was assayed. However, the enzyme

did not cleave any of these substrates (Fig. 4-32 C-E). Thus, our data speak

against cis-neurosporene or cis- lycopene cleavage by CCD4 as the origin of

an unidentified signaling molecule.

C

5 10 15 20 25 30 35 40

con

AtCCD4

proneurosporene 430 nm407

437 nmD

0 5 10 15 20 25 30 35 40

con

prolycopene

AtCCD4

mA

U 4

00 n

m

E

0 5 10 15 20 25

mA

U 4

50-5

00 n

m 9‘-cis-neurosporene

conAtCCD4

440 nm

469

416

min

min

Page 126: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

112

4.6.2 AtCCD7-mediated cleavage of acyclic poly-cis carotenes

AtCCD7 is well known to be specific for 9-cis-β-carotene cleaving at C9=C10.

This represents the initial step of strigolactone (SL) biosynthesis (Al-Babili and

Bouwmeester, 2015). The enzyme shows only minor activity with all-trans-β-

carotene and xanthophylls (Bruno et al., 2014). Its specificity towards acyclic

carotenes has not yet been investigated. Given the recurrent motif of 9-cis-

carotenoids being phytohormone precursors such as SLs and ABA (see above

and 1.2.3), AtCCD7 appears to be a prime candidate for the cleavage of

acyclic cis-carotenes of the desaturation pathway since all carry 9-cis-

configured double bonds (Fig. 4-33).

Consequently, we first assayed AtCCD7 with the four ζ-carotene isomers, with

9-cis-ζ-carotene amongst them (Fig. 4-33 A). Strikingly, 9-cis-ζ-carotene was

efficiently cleaved by AtCCD7 yielding the product P1 (Fig. 4-33 A)

accompanied by small amounts of additional products with

chromatographic and spectral similarity with P1 (Fig. 4-33 A, asterisks). Most

likely, they arise from P1 by non-enzymatic cis-trans isomerization upon sample

processing. In contrast, AtCCD7 did not convert any of the canonical

carotene desaturation intermediates 9,15,9’-tri-cis-, 9,9’-di-cis- and all-trans-ζ-

carotene (Fig. 4-33 A). Traces of product P1 upon incubation with these

isomers were most likely due to cross contamination with certain amounts of

the 9-cis isomer (Fig. 4-33 A, Ϯ). The sole exception is minor cleavage of all-

trans-ζ-carotene yielding product P1*, most likely the trans-isomer of P1.

Page 127: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

113

D

5 10 15 20 25 30

con

all-trans-lycopene

AtCCD7

10 20 30 40

con

9,9‘-di-cis-ζ-carotene

AtCCD7

P1Ϯ

con

9,15,9‘-tri-cis-ζ-carotene

AtCCD7

P1Ϯ

mA

U 4

00-7

00 n

m

con

9‘-cis-neurosporene

AtCCD7

P1

440 nm

469

416

B

5 10 15 20 25 30

mA

U 4

00-4

50 n

m

C

5 10 15 20 25 30

min

con

9-cis-lycopene

AtCCD7

P2P3

460 nm455 nm 466 nm

496440

360Ϯ

A

10 20 30 40

con

9-cis-ζ-carotene

AtCCD7

P1

* * *

428 nm

P1

0 10 20 30 40

min

10 20 30 40

P1 P1*con

all-trans-ζ-carotene

AtCCD7

0

400 nm424

379

398 nm

422377

295

400 nm

424

379

400 nm

424

379

471 nm

502445

Page 128: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

114

Fig. 4-33 HPLC analysis of AtCCD7 activity with canonical carotene desaturation

intermediates and 9-cis-isomers.

(A) Cleavage of ζ-carotene isomers by AtCCD7. Ϯ indicates 9-cis-ζ-carotene cross

contaminations. * indicates non-enzymatic isomerisation products of P1, including

P1*. (B – D) Cleavage of 9’-cis-neurosporene, 9-cis-lycopene and all-trans-lycopene

by AtCCD7, respectively. Ϯ in C indicates the appearance of the all-trans-isomer of

lycopene by spontaneous cis-to-trans isomerization. (E – G) Cleavage of phytofluene

isomers, proneurosporene and prolycopene by AtCCD7. UV/VIS spectra of the

substrates are given as insets or in Fig. 4-32. UV/VIS spectra of the products P1, P2 and

AtCCD7

AtCCD7

AtCCD7

mA

U35

0 nm

con

con

9-cis-phytofluene

9,15-di-cis-phytofluene

min0 10 20 30

all-trans-phytofluene

con

10 20 30

10 20 30

E

0 10 20 30 40

min

mA

U40

0-50

0 nm

prolycopene

con

AtCCD7

con

AtCCD7

proneurosporene

10 20 30 40

F

G

Page 129: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

115

P3 are given as insets. Note the relatively small product peaks in relation to substrate

decrease, caused by drastically lower extinction coefficients for the products

compared to the carotene substrates. HPLC system 9 was used for B, HPLC system 8

for C and D and HPLC system 7 was used for A, E, F and G. For structures of the

substrates, see Fig. 4-7.

This observation prompted us to assay other non-canonical 9-cis-configured

linear carotenes. Indeed, minor cleavage of 9’-cis-neurosporene, produced

from proneurosporene (7,9,9´-tri-cis-neurosporene) by CRTISO (Isaacson et al.,

2004; Yu et al., 2011) yielded product P1 (Fig. 4-33 B). Moreover, 9-cis-

lycopene was converted and products P2 and P3 were formed. They

displayed similar chromatographic and spectral characteristics and

presumably represent geometric isomers of each other arising from non-

enzymatic cis-trans isomerization (Fig. 4-33 C). Consistently, all-trans-lycopene

was not cleaved (Fig. 4-33 C) so that 9-cis specificity of AtCCD7 indeed

persists for linear carotenes. As observed for the canonical, poly-cis-

configured ζ-carotene isomers, AtCCD7 did not convert proneurosporene

(Fig. 4-33 F), prolycopene (7,9,7’,9’-tetra-cis-lycopene; Fig. 4-33 G) and 9,15-

di-cis-phytofluene as canonical carotene desaturation intermediates. In

addition, the non-canonical 9-cis-, 15-cis- and all-trans-isomer of phytofluene

were not recognized as substrates (Fig. 4-33 E).

In summary, AtCCD7 readily cleaves the non-canonical 9-cis isomers of ζ-

carotene and to a smaller extent of neurosporene and lycopene – whereas

the all-trans-species and the enzymatically formed poly-cis-configured

species of these carotenes are not cleaved.

4.6.3 Identification of apocarotenoids originating from acyclic 9-cis-

carotenes by AtCCD7 cleavage

The products derived from 9-cis-ζ-carotene (P1) and 9-cis-lycopene (P2, P3)

were collected from HPLC and subjected to LC-MS analysis for identification

Page 130: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

116

based on molecular masses and assuming that a canonical oxygenase

cleavage had occurred, leading to apocarotenoids with one incorporated

oxygen atom.

Fig. 4-34 Identification of the 9-cis-ζ-carotene cleavage product formed by AtCCD7

using LC-MS and GC-MS.

(A) LC-MS analysis of the HPLC-purified 9-cis-ζ-carotene cleavage product P1

including minor byproducts (see Fig. 4-33 A, asterisks). UV/VIS spectra are depicted in

insets. (B) GC-MS head space analysis of AtCCD7 assays with 9-cis-ζ-carotene and

an authentic geranylacetone standard. Spectral comparison was carried out with

the NIST 2.0 database. (C) Cleavage pattern of 9-cis-ζ-carotene by AtCCD7 and

structure of the cleavage products. For structures of the substrates, see Fig. 4-7.

O

geranylacetone

A

C

C9‘

9-cis-ζ-carotene

C10‘

B

6 8 10 12 14

min

Rel

ativ

e in

tens

ity 10.69

geranylacetone STD

10.67 m/z 69; 194

AtCCD7control

RT: 7.91

360 380 400

m/z

Rel

ativ

e A

bund

ance

C27H39O

379.2994RT: 7.86

360 380 400

m/z

Rel

ativ

e A

bund

ance

379.2993C27H39O

m/z 379.29954 (± 5 ppm)

7.6 7.8 8.0 8.2 8.4min

Rel

ativ

e A

bund

ance

7.91

7.86

419 nm412 nm

RT: 7.86 RT: 7.91

O

P1: 9-cis-ζ-apo-10‘-carotenal

C9

C10

C27H38O; m/z 378.29227

C13H22O; m/z 194.16707

60 80 100 120 140 160 180 200

m/z

69.11

151.13107.15 136.18

93.10 125.1683.1153.07

161.19 176.24194.28

Rel

ativ

e in

tens

ity

O

geranylacetone

C13H22O

AtCCD7

Page 131: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

117

Product P1 originating from 9-cis-ζ-carotene (Fig. 4-33 A) was analyzed. Two

product peaks with very similar chromatographic and spectral characteristics

as well as isobaric masses were detected by LC-MS suggesting their

spontaneous cis-trans interconversion upon sample processing (Fig. 4-34 A).

The molecular mass corresponded to C27H38O and P1 was therefore identified

as ζ-apo-10-carotenal (Fig. 4-34 A). Given the cleavage specificity of AtCCD7

for the C9-C10 double bond localized in the trans half-side of its 9-cis-

configured substrate (Bruno et al., 2014), the product P1 was identified as 9-

cis-ζ-apo-10’-carotenal (Fig. 4-34 C). Accordingly, P1* originating from minor

all-trans-ζ-carotene cleavage (Fig. 4-33 A) most likely represents the all-trans

species. To further corroborate these findings, we analyzed the head space of

AtCCD7 assays with 9-cis-ζ-carotene by GC-MS. As expected, we detected

and identified the second, volatile product arising from C9-C10 cleavage,

namely geranylacetone (Fig. 4-34 B). Taken together, AtCCD7 cleaved 9-cis-

ζ-carotene at C9-C10 to yield 9-cis-ζ-apo-10’-carotenal and geranylacetone

(Fig. 4-34 C). Accordingly, the minor cleavage activity with 9’-cis-

neurosporene yielded 9-cis-ζ-apo-10’-carotenal as well (Fig. 4-33 B).

Using the same experimental approach, the products P2 and P3 obtained

upon incubation of AtCDD7 with 9-cis-lycopene were identified. Product P2

was shown to co-elute with authentic all-trans-apo-10-lycopenal (Fig. 4-35 A).

The two products possessed isobaric masses corresponding to C27H36O,

confirming both cleavage products as apo-10-lycopenal (Fig. 4-35 B). Given

the cleavage specificity of AtCCD7 (Bruno et al., 2014), P3 was identified as 9-

cis-apo-10’-lycopenal. This is assumed to represent the main cleavage

product of AtCCD7 that readily isomerizes non-enzymatically to all-trans-apo-

10’-lycopenal (P2) under assay conditions and during sample processing. This

is corroborated by the observation that the 9-cis double bond in lycopene

readily isomerized to trans (Fig. 4-35 A, Fig. 4-33 C).

Page 132: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

118

Fig. 4-35 Identification of the 9-cis-lycopene cleavage product formed by AtCCD7

using LC-MS.

(A) The chromatographic behavior of the 9-cis-lycopene cleavage products P2 and

P3 (see Fig. 4-33 C for UV/VIS spectra) was compared to authentic all-trans-apo-10’-

lycopenal. HPLC system 8 was used for separation. (B) LC-MS analysis of HPLC-

purified P2 and P3. (C) Cleavage pattern of 9-cis-ζ-carotene by AtCCD7 and

structure of the cleavage products. For structures of the substrates, see Fig. 4-7.

Taken together, the results support that AtCCD7 cleaves 9-cis-lycopene at

C9-C10 yielding 9-cis-apo-10’-lycopenal and the volatile pseudoionone (Fig.

AtCCD7

min0 10 20 30

P2P3

9-cis-lycopene m

AU

400-

500n

m

all-trans-apo-10-

lycopenal

P2

P3

mA

U40

0-50

0nm

7 8 9 10min

376.5 377.0 377.5 378.0

m/z

Rel

ativ

e A

bund

ance

377.3531377.2839C27H37O

377.3530

apo-10-lycopenal STD

P2

P3

377.2835C27H37O

377.2836C27H37O

m/z 377.28389

A B

pseudoiononeC9‘

9-cis-lycopene

C10‘

P3: 9-cis-apo-10‘-lycopenal

C9

C10

C27H36O; m/z 376.27662

C13H20O; m/z 192.15142

[M+H]+

C

C9

C10

Page 133: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

119

4-35 C), although the latter could not be detected by GC-MS. This is in

accordance with C9-C10 cleavage of 9-cis-ζ-carotene and 9’-cis-

neurosporene as given above. Consequently, AtCCD7 maintained its regio-

specificity of cleavage and its 9-cis-specificity (Bruno et al., 2014) when

incubated with acyclic carotenes. Nevertheless, it needs to be emphasized

that cleavage of the 9’-cis-neurosporene and -lycopene was very weak

compared to 9-cis-ζ-carotene. The newly identified 9-cis-ζ-apo-10’-carotenal

and -lycopenal formed by AtCCD7 can be considered as potential

precursors of the postulated regulatory apocarotenoid(s) (Kachanovsky et al.,

2012; Avendaño-Vázquez et al., 2014).

4.6.4 Secondary cleavage of 9-cis-ζ-apo-10’-carotenal by AtCCDs

In order to exert signaling function, apocarotenoids need to attain increased

polarity to move between aqueous cell compartments. Since C27

apocarotenoids such as 9-cis-ζ-apo-10’-carotenal and 9-cis-apo-10’-

lycopenal are very hydrophopobic, further modification, e.g. cleavage by

CCDs, is required to give rise to a signaling molecule. This is exemplified by the

biosynthesis of strigolactones (see 1.2.3). The initial cleavage of 9-cis-β-

carotene by AtCCD7 to form 9-cis-β-apo-10’-carotenal is followed by further

cleavage, cyclization and oxygenation by AtCCD8 form carlactone, the

universal strigolactone precursor capable of leaving the plastid (Alder et al.,

2012) It is conceivable that 9-cis-ζ-apo-10’-carotenal and 9-cis-apo-10’-

lycopenal follow the same track being converted by AtCCD8 into

strigolactone-like compounds. Consequently, 9-cis-ζ-apo-10’-carotenal was

HPLC-purified from AtCCD7 assays and then subjected to AtCCD8 assays (Fig.

4-36). However, despite repeated approaches no conversion was achieved

(Fig. 4-36), whereas the strigolactone precursor 9-cis-β-apo-10’-carotenal in

the positive control was readily converted into carlactone (not shown).

Alternatively, further cleavage by AtCCD1 and AtCCD4, both known for their

ability to convert apocarotenoids, was taken into consideration. Amongst the

Page 134: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

120

two CCDs, only AtCCD1 cleaved the substrate as witnessed by the

disappearance of the substrate (Fig. 4-36). However, no obvious cleavage

product with an absorption in the wavelength range of 250 – 400 nm

appeared upon HPLC analysis. This is attributed to the known fact that CCD1

has broad substrate and regio-specificity regarding cleavage, leading to

complete substrate degradation (Walter and Strack, 2011).

Fig. 4-36 Cleavage of 9-cis-ζ-apo-10’-carotenal by AtCCD 1, 4 and 8.

The UV/VIS spectrum of the substrate is given as inset, for its structure see Fig. 4-34 C.

HPLC system 7 was used for separation and for purification of the apocarotenal.

The same approach was used with 9-cis-apo-10’-lycopenal. Given the minor

cleavage activity of AtCCD7 with 9-cis-lycopene, the yield of the

apolycopenal first needed to be optimized. Unfortunately – despite much

experimental variation – the yields remained too low for assaying the

apocarotenoids conversion by AtCCDs.

5 10 15 20 25 30 350 40

mA

U42

8 n

m

con

AtCCD1AtCCD4

AtCCD8

9-cis-ζ-apo-10‘-carotenal

428 nm

min

Page 135: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Results

121

Fig. 4-37 Structures of acyclic carotene substrates tested for AtCCD4 and AtCCD7.

(I), 9,15-di-cis-phytofluene, (II) 9-cis-phytofluene, (III) all-trans-phytofluene, (IV) 9,15,9’-

tri-cis-ζ-carotene, (V) 9,9’-di-cis-ζ-carotene, (VI) 9-cis-ζ-carotene, (VII) all-trans-ζ-

carotene, (VIII) 7,9,9’-tri-cis-neurosporene, (IX) 9’-cis-neurosporene, (X) 7,9,7’,9’-tetra-

cis-lycopene, (XI) all-trans-lycopene, (XII) 9-cis-lycopene. Scissors indicate cleavage

positions at the C9-10 or C9’-10’ double bonds for AtCCD7, scissors in light grey

indicate minor cleavage activity of AtCCD7. AtCCD4 did not cleave any of the

substrates.

II

9

10

VII

10‘IV

9‘

15 15‘

9 10

V

9‘

10‘910

VIII

910 9‘

10‘

78

XI

9

10

XII

X 7‘

99‘

7

VI

910

CCD7

CCD7

I

910

1515‘

III

9‘IX

10‘CCD7

Page 136: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

122

5 Discussion

5.1 Phytoene desaturase PDS: evaluation of implications deduced

from the OsPDS-His6 structure

The crystal structure of OsPDS-His6 has been resolved at a 2.77 Å resolution (in

collaboration with the team of Prof. Dr. Einsle, Department of Chemistry,

University of Freiburg) and was published recently (Brausemann et al., 2017).

However, resolved structures often raise questions on function which can only

be addressed by drawing educated conclusions in the absence of functional

investigations. It is one incentive of this thesis to complement the implications

made by providing experimental data.

5.1.1 OsPDS-His6 is a monotopic membrane protein

The OsPDS-His6 structure revealed a membrane binding domain consisting of

an uncharged surface area around the substrate cavity entrance,

surrounded by a ring of positively charged residues compensating for the

negative charge of phospholipids. These features are typical of monotopic

membrane proteins (Brausemann et al., 2017 and references therein). In line

with this, OsPDS-His6 could be purified as soluble in the absence of detergents

but it attached spontaneously to phosphatidylcholine liposomes. This was

found to be independent of the presence of the membrane-soluble

substrates phytoene and DPQ (Fig. 4-1). OsPDS-His6 could not be detached

from the proteoliposomes by high salt treatment (Fig. 4-1); membrane

solubilization by detergents was required for this purpose. Thus, the uncharged

surface area interacts hydrophobically with the membrane core. Freeze

fracture electron microscopy of proteoliposomes revealed no sign of

membrane spans (Fig. 4-2); their absence is also supported by hydropathy

plots. The protein is attached to the membrane surface, interacting with only

one lipid layer.

Page 137: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

123

5.1.2 OsPDS-His6 employs an unprecedented “flavin only” reaction

mechanism during carotene desaturation

The active site of OsPDS-His6, i.e. the portion of the substrate cavity that

surrounds the redox-active isoalloxazine moiety of FAD in a 8 Å radius,

contains only few charged residues (Fig. 4-5 A). Only Arg300 might be in an

appropriate distance (~ 4 Å) to potentially make a functional contribution.

This lack of charged residues is unprecedented in flavin dehydrogenases

implying that PDS, unlike the inhomologous bacterial phytoene desaturase

CrtI (Schaub et al., 2012), does not employ an acid-base component in its

carotene desaturation reaction mechanism. Accordingly, and in contrast to

CrtI, no deuterium incorporation into carotene desaturation products could

be observed in a D2O environment (Fig. 4-6; Schaub et al., 2012). Moreover,

Arg300 to Ser mutagenesis demonstrated that the enzyme retained enzymatic

activity at lowered levels. This shows that Arg300 is not catalytically essential.

Accordingly, Hydrilla PDS maintains substantial activity when the Arg300

homolog is substituted by virtually any other proteinogenous amino acid

(Arias et al., 2006). The reduced activity of the mutant might however

indicate an ancillary function of Arg300 such as polarizing the carotene π

electron system, thereby facilitating desaturation of the adjacent C11-C12

site. Thus, two “flavin only” reaction mechanisms in which the redox-reactive

N1-N5 functionalities of the isoalloxazine act as sole catalysts conceivably

carry out the trans-desaturation of phytoene at position C11-C12 and trans-

to-cis isomerization at position C9-C10 (Fig. 4-5; Brausemann et al., 2017).

5.1.3 OsPDS-His6 follows an ordered ping pong bi bi mechanism

OsPDS-His6 is equipped with a single exceptionally long hydrophobic substrate

cavity (~ 43 Å). Modeling suggests (Fig. 4-3) that it can entirely accommodate

the carotene substrates as well as the co-substrate PQ, but it cannot bind

them simultaneously. This implies an ordered ping pong bi bi mechanism: PDS

desaturates phytoene, expels phytofluene and then accommodate the PQ

Page 138: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

124

required for reoxidation of the FADred formed. The PQH2 formed leaves the

cavity to allow another round of carotene desaturation.

The findings presented are compatible with this notion. First, one round of

phytoene desaturation occurred in the absence of PQ indicating that

simultaneous binding of both substrates is not required (see 4.1.4).

Phosphatidylcholine from a natural source (soybean) was used in this

experiment so that quinone impurities cannot be excluded. However, using

TritonX100 for substrate solubilization instead of the lipid gave the same result.

Second, the intermediate phytofluene is detectable and, thus, in fact

released from OsPDS-His6 into liposomal membranes. The third indication

comes from NFZ inhibition. NFZ presumably occupies the PQ binding site within

the substrate cavity that accommodates phytoene as well. Nevertheless, NFZ

was found to only compete with DPQ but not with phytoene (Fig. 4-19), an

observation that is explicable by the sequential binding of the two substrates.

Given an ordered ping pong bi bi mechanism, the thermodynamics of

carotene desaturation and PQ reduction come into question. The findings

presented support that PQ is only required for FADred reoxidation to allow

repeated desaturation cycles. The independence of both partial reactions,

namely carotene desaturation and PQ reduction, demonstrates that both

represent exergonic processes. This contradicts the long standing hypothesis

that the endergonic process of phytoene desaturation requires coupling to

the photosynthetic redox chain in chloroplasts or to a respiratory redox chain

in chromoplasts as exergonic processes involving PQ (see 1.3.1.1). The

exergonic character of the “flavin only” mechanism is corroborated by the

observations that the equilibrium of the OsPDS-His6 reaction is far on the side

of the end product ζ-carotene (see 4.2.1). Thus, the dependence of PDS

activity on the plastid PQ redox state (Nievelstein et al., 1995) should rather be

viewed in the context of a kinetic control PQ exerts on FADred reoxidation that

is required for the progression of the overall reaction.

Page 139: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

125

It remains to be elucidated how the ordered ping pong bi bi mechanism is

achieved, i.e. how preferential carotene substrate binding prior to PQ binding

is orchestrated. It is conceivable that the FAD redox state acts as a switch

triggering conformational changes between a fold that is prone to carotene

binding (FADox) and one that preferentially binds PQ (FADred). If true, the

structure of OsPDS-His6 in complex with NFZ would represent the reduced

enzyme state because this redox-inactive inhibitor prevents FADred

reoxidation, in contrast to its structural analogue PQ (Gemmecker et al.,

2015), Testing this hypothesis would require resolving the structure of oxidized

OsPDS-His6. However, this proved to be impossible since NFZ acted as an

indispensable protein stabilizing agent (Gemmecker et al., 2015).

It is worth noting that the bacterial phytoene desaturase CrtI most likely differs

from PDS regarding the ordered ping pong bi bi mechanism. CrtI introduces

four double bonds into 15-cis-phytoene, two on each half side. During the 4-

step carotene desaturation, FADred reoxidation is mediated by reduction of

oxygen that is readily accessible to the active center while the carotene is still

bound (Schaub et al., 2012). Thus, there is no large lipophilic co-substrate

requiring displacement. Moreover, contrary to expectation for an ordered

ping pong bi bi mechanism, CrtI barely releases intermediates (Schaub et al.,

2012). CrtI cannot desaturate carotenes in the absence of oxygen (Schaub et

al., 2012). Therefore, it is likely that oxygen is required in the active site during

carotene desaturation to provide the necessary change in free energy.

5.1.4 Substrate and regio-specificity of PDS: role of the 15-cis-configuration

PDS exhibits high regio-specificity as witnessed by the fact that no carotene

desaturation at a position other than C11-C12 and C11’-C12’ or isomerization

at a position other than C9-C10 and C9’-C10’ has ever been reported.

Accordingly, 9,15-di-cis-phytofluene and 9,15,9‘-tri-cis-ζ-carotene are the sole

products. The 15-cis-isomer of phytoene is long known as the PDS substrate. It

was shown to be converted 10 x faster than the all-trans isomer (Breitenbach

Page 140: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

126

et al., 2005). Notably, desaturation of all-trans-phytoene yielded other ζ-

carotene isomers than the tri-cis isomer. This implies a relevance of 15-cis-

configuration in terms of substrate recognition and regio-specificity.

Regio-specificity relies on the correct positioning of the correct carotene

single bond relative to the catalytic isoalloxazine of FAD. It is an interesting

note in this context that a 9-cis-configured double bond arrests the

xanthophyll substrate in the 9-cis-epoxy-carotenoid cleavage dioxygenase

VP14 in the correct position (Messing et al., 2010). The 15-cis double bond of

phytoene might function analogously in PDS. Alternatively, the carotenoid

substrate might be introduced until reaching the cavity end, i.e. it is the

substrate length that might mediate substrate positioning. The modeling of 15-

cis-phytoene in its extended conformation into the substrate cavity of OsPDS-

His6 (Brausemann et al., 2017) does not allow distinguishing between these

alternatives. The kinked structure of the cavity appears suitable to

accommodate the kinked central C15-C15’ double bond but the

contribution of the substrate length appears just as likely (Fig. 4-3; see 4.1.2).

To disentangle these alternatives, a C5-truncated variant of 15-cis-phytoene,

15-cis-1’,2’,3’,16’,17’-penta-nor-phytoene (hereafter 15-cis-nor-phytoene; Fig.

4-4 A) was used as substrate. With this asymmetrically shortened substrate,

regio-specificity would be maintained at C11-C12 and C11’-C12’ if the 15-cis

double bond rather than molecule length was decisive (see 4.1.2). In fact, LC-

MS analysis revealed that 15-cis-nor-phytoene was desaturated twice to yield

an end product with a UV/VIS spectrum resembling the one of 9,15,9-tri-cis-ζ-

carotene suggesting that desaturation and isomerization had occurred at the

correct sites (Fig. 4-4 B). Thus, the 15-cis double bond in phytoene, and

potentially the entire triene chromophore region of which it forms part,

mediates regio-specificity. This is supported by the fact that elongation of the

cavity back end by site-directed mutagenesis, i.e. by the mutations Tyr506Phe

and Thr508Val, did not affect regio-specificity (see 4.1.2.2).

Page 141: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

127

With 15-cis-nor-phytoene, the intermediate – being desaturated only once (at

one end) – predominated quantitatively (Fig. 4-4 B). Conceivably, the

truncated half side of 15-cis-nor-phytoene is bound less efficiently by the

enzyme, i.e. the length of the molecule half sides might be decisive for

binding without affecting regio-specificity. Thus, although PDS

accommodates its C40 substrate entirely within the cavity, there appears to

be a sidedness of recognition (“half side recognition”) of the symmetric 15-cis-

phytoene that also applies to an asymmetric, truncated carotene. In line with

this, 9,15-di-cis-phytofluene – of which one half side is identical to a phytoene

half side – is as readily converted as 15-cis-phytoene (see 5.1.5).

The triene chromophore in phytoene might contribute substantially to attain

regio-specificity. Its role should be further investigated by assessing whether

regio-specificity is maintained with a synthetic 15-cis-phytoene derivative

lacking the double bonds at C13-C14 and C13’-C14’ or with synthetic 15-cis-

phytofluene. Moreover, non-covalent interactions between the chromophore

region and PDS residues could be deduced from computation according to

the density functional theory (Johnson et al., 2010), provided in silico docking

of 15-cis-phytoene (or truncations thereof) turns out to be feasible.

It is worth noting that PDS and the inhomologous bacterial phytoene

desaturase CrtI differ regarding substrate cavity topology, mechanisms of

substrate recognition and regio-specificity, despite similar tertiary structures

(Fig. 4-24). In silico docking carried out with the CrtI apoprotein revealed that

the cavity only accommodates C18 – C20 truncations of carotenes, i.e.

carotene half sides (Schaub et al., 2012). This implies dynamic CrtI

dimerization upon binding of a phytoene half side, with the dimer sequentially

introducing all four double bonds, minimizing intermediate release. Notably,

as with PDS, substrate molecule length is not decisive for regio-specificity of

carotene desaturation. It was maintained with C30 15-cis-

(1,2,3,16,17,1’,2’,3’,16’,17’)-deca-nor-phytoene, corresponding to phytoene

truncated by C5 on each end (Raisig and Sandmann, 2001).

Page 142: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

128

5.1.5 Kinetic characterization and mathematical modeling of OsPDS-His6

5.1.5.1 Modeling supports substrate channeling in OsPDS-His6 homotetramers

GPC, negative staining electron microscopy and protein crystallography of

soluble OsPDS-His6 suggested the enzyme to be flavinylated and active as

homotetramer (Fig. 1-3 B); monomers were found to be deflavinylated and

irreversibly inactivated (Gemmecker et al., 2015). During this thesis, freeze

fracture electron microscopy with OsPDS-His6 proteoliposomes revealed a PDS

particle size that supports the homotetramer as the enzyme’s active state at

membrane surfaces (Fig. 4-2). The ordered ping pong bi mechanism supports

that each subunit within the homotetramer can act independently (see 4.1.4

and 5.1.3). This raises questions on the mechanistic relevance of

homooligomerization.

In a consequence of the ordered ping pong bi bi mechanism, PDS represents

a bifunctional phytoene-phytofluene desaturase catalyzing a two-step

enzyme cascade (McClure et al., 1969). Each phytofluene formed from

phytoene must be released and both carotene substrates compete for

binding. Depending on the relative concentrations of substrate and

intermediate, this mechanistic requirement inherently counteracts formation

of the end product ζ-carotene – if no mechanistic “countermeasures” are

present to favor rebinding of phytofluene over phytoene binding.

Homooligomerization can represent such a “countermeasure” paving the

way for cooperativity in substrate binding or channeling of the intermediate

between subunits. Enhanced affinity for carotenes or restricted intermediate

diffusion, respectively, would facilitate phytofluene rebinding and conversion

by a different PDS subunit containing oxidized FAD. However, substrate

concentration dependencies for phytoene or phytofluene did not reveal

sigmoidality (Fig. 4-14), a hallmark of cooperativity (Copeland, 2000).

Substrate channeling has predominantly been reported for heterooligomeric

complexes (Lee et al., 2012) but its occurrence in PDS homotetramers

Page 143: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

129

appears likely, given the extraordinary situation with a symmetric substrate

undergoing two formally identical reactions at half sides (Fig. 1-3).

Substrate channeling between a “source” enzyme and a “sink” enzyme

requires substrate cavity entrances to be within less than 10 Å distance and

limited intermediate diffusion to attain locally enhanced intermediate

concentrations (Wheeldon et al., 2016; Lee et al., 2012). Both prerequisites

appear to be met by the PDS homotetramer, with its subunits representing the

“source” as well as the “sink” for the intermediate. Four cavity entrances point

towards each other (Fig. 1-3 B) and it appears likely that the monotopic

interaction with lipids creates a membrane microdomain restricting diffusion

thereby favoring the enrichment of nascent phytofluene in proximity to the

PDS homotetramer (Lee et al., 2012). In this regard, higher order

oligomerization may provide a larger “sink” than dimerization, the minimal

association required for channeling.

Being inaccessible to direct observation, we resorted to dynamic

mathematical modeling in order to investigate the occurrence of substrate

channeling. All models examined encompassed phytoene desaturation,

phytofluene desaturation and PQ reduction (for details, see 4.2). Briefly, the

“monomeric model” (Fig. 4-8 A) that is based on independently acting

subunits within homotetramers could not describe PDS behavior, in contrast to

the “substrate channeling model” (Fig. 4-8 B) that implements substrate

channeling in homotetramers. This means that substrate channeling is a

necessary process to describe the PDS behavior. This model supports two

kinetically inequivalent fates of phytofluene. The nascent intermediate is

released into a membrane microdomain that limits its diffusion. It is then either

channeled into a different oxidized subunit of the same homotetramer for

rapid conversion or diffuses into the membrane space. Here, it can diffuse

without any limitation to eventually be taken up and converted to ζ-carotene

by any oxidized subunit. The rate constants deduced from the model (Table

4) support this hypothesis suggesting channeled phytofluene to be converted

Page 144: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

130

5 x faster than phytofluene that diffuses without limitation within the

membrane. Computation of the carotene fluxes revealed that nascent

phytofluene has a relatively low probability of leaving the microdomain,

indicating efficient limitation of diffusion.

However, substrate channeling is not perfect. It is susceptible to disturbances,

because such microdomains are in equilibrium with the membrane. A first

indication is provided by the considerable delay between the inset of

phytofluene and ζ-carotene formation during PDS reaction time courses and

the considerable level of released phytofluene (Fig. 4-7 D; Wheeldon et al.,

2016). A second indication comes from the observation that excess phytoene

supply favors phytofluene formation relative to ζ-carotene formation (Fig. 4-

14) because phytoene progressively outcompetes (nascent) phytofluene in

microdomains in terms of enzyme binding. Accordingly, modeling supports

that excess phytoene supply, i.e. high carotene flux through PDS, impedes the

formation of the end product ζ-carotene.

5.1.5.2 Rate-limiting steps in OsPDS-His6

The substrate concentration-dependent behavior of OsPDS-His6 was

investigated both experimentally to elucidate Michaelis-Menten parameters

and by simulation using the established mathematical model (see 4.2.3).

It is to be mentioned that the Michaelis-Menten parameters determined need

to be treated with some caution for the two-step reactions of PDS and ZDS.

The two carotene desaturation steps represent separate events with the

velocity of intermediate (phytofluene and neurosporene, respectively)

desaturation depending on the velocity of substrate (phytoene and ζ-

carotene, respectively) desaturation. Thus, the Michaelis-Menten parameters

deduced from the velocity of the overall reaction (Table 5) – i.e. formation of

the end product from the initial substrate – represent composite, “apparent”

values encompassing the velocities of both partial reactions. This is not the

case when the intermediate (phytofluene and neurosporene, respectively)

serve as substrates. Thus, in order to obtain an appropriate measure of affinity

Page 145: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

131

for the initial substrate (phytoene in PDS, ζ-carotene in ZDS), Michaelis-Menten

parameters should be deduced from the velocity of intermediate formation

rather than deducing the parameters from the overall 2-step reaction.

Accordingly, phytofluene formation from phytoene could not be saturated

and no KM could be determined (Fig. 4-14 B). In contrast, a KM of 66.8 ± 20.7

mM was determined for phytofluene as substrate. Consequently, the KM for

phytoene is far higher than for phytofluene. Interpreting KM in terms of affinity,

the intermediate is preferred over the substrate. This is reflected by the rate

constants determined by mathematical modeling, implying that phytofluene

deposited in liposomes is converted twice as fast as phytoene deposited in

liposomes (Table 4). This would in fact reflect different affinities because the

desaturation process is formally identical for phytoene and phytofluene. Thus,

higher affinity for the intermediate than for the initial substrate appears as

another feature to favor end product formation over intermediate release.

Applying the same rationale, OsPDS-His6 exhibits a by far higher affinity for

DPQ than for any of the carotene substrates with a KM of only 1.3 ± 0.2 mM. In

line with this, the mathematical model indicates that FADred reoxidation by

DPQ is much faster than any carotene desaturation process (Table 4, Fig. 4-12

E), ensuring that OsPDS-His6 is rapidly reoxidized (Fig. 4-12 D). This suggests that

FADred reoxidation is not rate-limiting for carotene fluxes through PDS, as long

as sufficient oxidized PQ is supplied (experiments were carried out under DPQ

saturation). When present in limiting amounts (Fig. 4-14 A), repeated carotene

desaturation cycles and carotene flux from phytoene via phytofluene to ζ-

carotene are hindered. Since ζ-carotene formation depends on previous

phytofluene formation, PQ deficiency favors intermediate formation relative

to end product formation (Fig. 4-14 A). Thus, higher affinity for PQ than for

carotene substrates can be considered as a third mechanism in PDS to favor

end product formation. It is conceivable that rapid reoxidation of the subunits

within homotetramers assists substrate channeling of the nascent phytofluene

that still resides within the PDS microdomain.

Page 146: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

132

The substrate channeling model and the rate constants deduced from

reaction time courses could in principle simulate and reflect the observed

substrate concentration-dependent relations of intermediate and end

product formation by OsPDS-His6 (Fig. 4-14). This supports the validity of the

model. It needs to be stated that it also systematically overestimates

Michaelis-Menten parameters, used as a quantitative measure to allow

comparisons, by factors of 1.1 to 4.1 (Table 5). The error is likely due to gradual

alterations of the structure and physicochemical properties of liposomal

membranes when increasing concentrations of the hydrocarbon substrates

are incorporated that conceivably impair OsPDS-His6 activity (Sikkema et al.,

1993). The model was established at low substrate concentrations and cannot

consider this structural circumstance upon extrapolation. Moreover, the

production of enzyme and liposomes batches with identical specific activities

showed to be notoriously difficult. The preparations used in reaction time

course experiments to develop the model were different from the ones used

to investigate substrate concentration dependencies. This contributes to the

quantitative deviations from the model despite qualitative similarities, too.

5.1.6 Characterization of norflurazon - OsPDS-His6 interactions

5.1.6.1 Quinone-competitive inhibition by norflurazon is mediated by several

non-covalent interactions

Phytoene desaturase is a prominent target for a structurally heterogeneous

group of bleaching herbicides such as norflurazon (NFZ), fluridone and

diflufenican sharing a meta-trifluoromethylphenyl (m-CF3-phenyl) moiety often

fused to a second (heterocyclic) carbonyl-substituted ring structure (see

1.3.1.3). The lack of structural information has long hampered a detailed

molecular characterization of the inhibitory mode of these herbicidal

compounds. With emerging resistances of weeds against commonly used

herbicides, knowledge of structure-function relationship can pave the way

towards the rational design of optimized herbicides and the development of

herbicide resistant PDS versions in crops. The importance of such investigations

Page 147: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

133

is reflected by the fact that they have met the interest of the agrochemical

company Syngenta and collaboration has been initiated. In accordance

with the corresponding confidentiality agreement, not all results can be

presented at this point. The refinement of the OsPDS-His6 structure in a

complex with NFZ was key to the investigations concerning PDS-targeting

herbicides (Brausemann et al., 2017).

The data presented reveal that NFZ exhibits competitive inhibition towards PQ

(Fig. 4-15) which is in accordance with results obtained with complex cell-free

assay systems (Breitenbach et al., 2001). Moreover, as shown in a preceding

publication (Gemmecker et al., 2015), NFZ is tightly bound to OsPDS-His6

during enzyme purification but can be displaced by DPQ. This suits the

structure-derived notion that PQ and NFZ share the same binding site. In

contrast, NFZ exhibits either uncompetitive or non-competitive inhibition

towards phytoene (Fig. 4-15), occupying the same substrate cavity. Non-

competitive inhibition is supported by the fact that the crystallized PDS-NFZ

complex represents an enzyme-inhibitor complex (Brausemann et al., 2017)

with no substrate bound. This situation can only occur for competitive or non-

competitive inhibitors whereas uncompetitive inhibitors sensu stricto require

the presence of the enzyme-substrate complex (Copeland, 2000). Moreover,

non-competitive inhibition towards phytoene has been suggested earlier

(Sandmann et al., 1989; Mayer et al., 1989).

This differential inhibition mode regarding the two substrates further supports

the prevalence of an ordered ping pong bi bi mechanism (see 5.1.3), with

phytoene or phytofluene binding to the FADox state of the enzyme and DPQ

and NFZ binding to the FADred state. In line with this, Gemmecker et al. (2015)

reported that NFZ binds and – being redox-inactive - stabilizes the reduced

enzyme state after carotene desaturation.

NFZ as quinone-competitive inhibitor lowers the apparent PQ availability,

thereby preventing FADred reoxidation and the execution of continuous

desaturation cycles (see 5.1.5.2). Thus, NFZ inhibition diminishes enzymatic

Page 148: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

134

activity, i.e. carotene flux from phytoene via phytofluene to ζ-carotene. This

favors phytofluene formation relative to ζ-carotene formation, as the latter

depends on the former in this two-step reaction. This fits with the observation

that low PQ supply increases the phytofluene proportion (Fig. 4-14). Previously,

this circumstance has been misinterpreted in terms of a differential NFZ

susceptibility of the two desaturation reactions.

Despite significant structural differences compared to NFZ, the m-CF3-phenyl

containing herbicides diflufenican and fluridone compete with PQ binding, as

shown here (Fig. 4-15) and as hypothesized earlier (Laber et al., 1999). This

emphasizes the importance of the common structural features, namely the

m-CF3-phenyl moiety and the presence of a carbonyl group in a second

rather hydrophobic ring structure. However, in silico docking of diflufenican

and fluridone into the PQ binding site was not successful, suggesting an

induced fit upon inhibitor binding. Further investigation would require

crystallography of OsPDS-His6 containing complexed diflufenican or fluridone.

These structures would provide valuable models for the calculation and

prediction of non-covalent interactions between PDS and inhibitors using the

density functional theory-based method of Johnson et al. (2010).

Such calculations were used during this thesis to elucidate the molecular basis

of the long-known importance of the m-CF3 substituent in NFZ (Mayer et al.,

1989; Sandmann et al., 1989; Sandmann and Böger, 1982), fluridone

(Sandmann et al., 1992) and diflufenican (Cramp et al., 1987). The role of this

functionality has mainly been attributed to its lipophilicity and

electronegativity so far. The quantum mechanical calculations presented

reveal that the m-CF3 substituent provides five weakly attractive non-covalent

interactions with hydrogen or sulphur atoms of Met188, Met277, Ala280 and

Phe423 (Fig. 4-16), all representing conserved PDS residues of the PQ binding

site (Fig. 4-20). Far less attractive interactions are expected when CF3 is

replaced by CH3, Cl or H and thus, the “magic” of CF3 indeed involves a

combination of lipophilicity, electronegativity and fluorine effects rather than

Page 149: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

135

a single “all-or-none” interaction. Such computational approaches when

applied to additional compounds could substantially facilitate the rational

design of new PDS-targeting herbicides – as opposed to cumbersome

combinatorial chemistry approaches.

Interestingly, despite the lack of structural information, early studies have led

to a fairly accurate prediction of the binding site of m-CF3-phenyl herbicides

by Laber et al. (1999). Indeed, the m-CF3-phenyl moiety is surrounded by a

conserved pocket of lipophilic residues such as Met188, Met277, Ala280, Phe423,

Phe162 and Met310 (Fig. 4-16). The conserved Arg300 represents the postulated

residue for hydrogen bonding of the oxygen functionality in vicinity to the m-

CF3-phenyl group.

5.1.6.2 Site-directed mutagenesis confers norflurazon resistance at the

expense of catalytic activity

The effective use of PDS-targeting herbicides in agriculture would benefit from

resistant crops. This can nowadays be achieved by applying state of the art

genome-editing methods since PDS mutations conferring resistance to NFZ

and fluridone have been identified in cyanobacteria, algae and in one plant

(see 1.3.1.3). However, it remained elusive how these mutations confer

resistance and in which way enzyme characteristics might be affected.

In order to address these questions, such PDS mutant enzymes were

generated and biochemically characterized (see 4.3.2). In summary,

individual mutation of the residues Phe162, Arg300 and Leu538 resulted in

enzymes that suffered a substantial loss of enzymatic activity in vitro

compared to the wild type enzyme, despite retaining a native-like tertiary

structure. In depth analysis of the most active mutant enzyme, Arg300Ser PDS,

revealed that homooligomerization was not affected either (Fig. 4-18).

However, the kinetic parameters were drastically altered (Table 6) as

witnessed by higher affinity for DPQ (KM lower by factor 3) and phytoene (KM

lower by factor 12) and lower Vmax regarding both substrates, the latter

reflecting the enzyme’s loss of activity. Moreover, as expected from reports

Page 150: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

136

about this mutation (Arias et al., 2006; Martinez-Férez et al., 1994), the affinity

for NFZ was found to be lowered as witnessed by a 7 fold higher Ki (Table 6).

Thus, the point mutation Arg300Ser led to global kinetic changes.

Removing a formal charge from an active center as in Arg300Ser PDS

represents a major change with possible long-range conformational changes

exerting multiple effects. More specifically, the increased active center

lipophilicity in the mutant might result in enhanced affinity for the lipophilic

substrates phytoene and DPQ (lower KM for both) and the decreased affinity

for the smaller, more hydrophilic inhibitor NFZ (higher Ki). These parameter

changes contradict the notion of a simple analogy in NFZ and DPQ binding

via hydrogen bonding of their keto functions (see 5.1.6.1) but plead for a

contribution of the hydrocarbon moiety in the case of PQ. Interpreting

lowered KM in terms of affinities, the increased affinity for substrates might as

well be accompanied by increased affinity for the similarly lipophilic products.

It may therefore be conceivable that product release and the rapid

sequential order of carotene and PQ binding are hindered in the mutant,

leading to diminished catalytic activity. In terms of herbicide resistance, the

relative affinities for DPQ and the quinone-competitive NFZ need to be

considered. Despite increased DPQ affinity, the enzyme exhibits increased

NFZ resistance (higher Ki). This emphasizes the importance of hydrogen

bonding of NFZ by Arg300 – besides other non-covalent interactions, e.g. with

the m-CF3-phenyl moiety (see 4.3.3). This supports the above given notion that

affinity for the more hydrophobic PQ does not solely rely on hydrogen

bonding of its keto function but strongly relies on additional non-covalent

interactions with its hydrocarbon moiety.

Substrate concentration dependencies for Arg300Ser PDS showed that

increasing phytoene concentrations favored the formation of the

intermediate phytofluene in relation to the end product ζ-carotene, like with

wild type OsPDS-His6 (compare Fig. 4-19 and Fig. 4-14). However, the mutant

preferentially released phytofluene and not ζ-carotene at any given

Page 151: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

137

phytoene concentration – in contrast to the wild type enzyme (compare Fig.

4-19 and Fig. 4-14). This was mainly due to strongly impaired ζ-carotene

formation, while phytofluene was present in wild type-like quantities. In the

light of the mathematical model for OsPDS-His6, the simplest explanation

would be that the lowered enzymatic activity and carotene flux through the

mutant enzyme mainly diminishes ζ-carotene formation as it additionally

depends on previous phytofluene synthesis. Mathematical modeling of

reaction time courses with Arg300Ser PDS was tried but did not allow further

disentangling the sub-processes affected in the mutant. Impaired substrate

channeling appears unlikely as a cause since homooligomerization as the

prerequisite remained unaffected (Fig. 4-18).

As a conclusion, all PDS mutant enzymes generated to confer NFZ resistance

(Arg300Ser, Arg300Thr, Leu538Phe, Leu538Arg and Phe162Val) in principle allowed

for ζ-carotene formation in E. coli, engineered to produce 15-cis-phytoene,

and in vitro – but at much lower rates than the wild type enzyme. In

consideration of the results obtained for Arg300Ser PDS, engineering NFZ

resistance trades-in lowered catalytic activity of the enzyme. However, this

might suffice for carotenogenesis in vivo, as witnessed by the wild type-like

carotenoid levels of NFZ-resistant mutants (see 1.3.1.3). It is conceivable that

the loss of activity is compensated by regulatory mechanisms in planta.

Genome editing would be the method of choice to investigate whether

engineering of PDS can produce agriculturally useful herbicide-tolerant plants

with uncompromised performance.

Page 152: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

138

5.2 ζ-carotene desaturase ZDS: a comparison to PDS

5.2.1 OsZDS-His6 utilizes an ordered ping pong bi bi mechanism and a

“flavin only” mechanism

In recent years and during this thesis, substantial progress has been made in

the understanding of PDS (Gemmecker et al., 2015; Brausemann et al., 2017;

see 4.1, 4.2, 4.3, 5.1). Little is known about ZDS, although many similarities are

anticipated because of homology and a similar function. Thus, investigations

on ZDS were initiated as this enzyme is crucial to an understanding of the

poly-cis pathway of carotene desaturation (for details, see 4.4).

OsZDS-His6 was natively purified in the absence of detergents to near

homogeneity as a yellow enzyme. LC-MS analysis revealed FAD as sole

cofactor bound by a conserved Rossman fold (Fig. 4-22). OsZDS-His6 rapidly

attached to substrate containing liposomes and is active in the absence of

supplemented FAD, yielding proneurosporene (PN) and prolycopene (PL)

from 9,9’-di-cis-ζ-carotene and exhibiting high regio-specificity (Fig. 4-21). Thus,

it represents a monotopic membrane flavoprotein. Its redox cofactor FAD

requires reoxidation by PQ, the latter serving as diffusible, terminal electron

acceptor. The occurrence of carotene desaturation in the absence of PQ

and the observed release of the intermediate from the enzyme revealed that

both carotene desaturation reactions and PQ reduction are per se

independent of each other (see 4.4.1.3). Consequently, OsZDS-His6 is

expected to follow an ordered ping pong bi bi mechanism and represents a

bifunctional ζ-carotene – neurosporene desaturase, catalyzing a two step-

desaturation cascade with formally identical reactions on the substrate half

sides. PQ is only required for FADred reoxidation to allow repeated reaction

cycles and exerts kinetic control over the overall reaction towards

prolycopene. Thus, regarding these basic mechanisms, similarities with PDS

are very significant.

In line with this, homology modeling (carried out because crystallization

attempts failed) revealed high structural similarity with OsPDS (Fig. 4-24). A

Page 153: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

139

single hydrophobic substrate cavity was present, capable of entirely

accommodating both carotene substrates and PQ but not simultaneously.

The FAD-containing active center of OsZDS was found to lack acidic and

basic amino acids that could mediate acid-base catalysis of carotene

desaturation. Thus, ZDS seems to employ a “flavin only” mechanism with the

isoalloxazine as the sole catalyst, similar to the reaction mechanism proposed

for PDS (for details, see 4.4.2).

Characterization of the substrate concentration-dependent kinetics of OsZDS-

His6 revealed high similarities with OsPDS-His6, supporting that ZDS utilizes similar

mechanisms to favor end product formation in its two-step carotene

desaturation. First, OSZDS-His6 exhibits higher affinity for the intermediate PN

than for the initial substrate ζ-carotene. The KM for PN is 13.9 ± 2.5 mM and,

following the rationale detailed for PDS that KM for the substrate should be

determined via formation of the intermediate (see 5.1.5.2), KM could not be

determined experimentally for ζ-carotene and is higher than 50 mM (see

4.4.3.3). Accordingly, increasing concentrations of the initial substrate ζ-

carotene favored formation of the intermediate over the end product, as it

competes with PN (Fig. 4-28 A). Thus, low carotene fluxes through ZDS are

essential to favor end product formation. Second, OSZDS-His6 exhibits higher

affinity for PQ than for its carotene substrates with a KM of only 0.3 ± 0.1 mM for

DPQ. This ensures rapid enzyme reoxidation for repeated carotene

desaturation cycles and conceivably facilitates rapid conversion of the

recently expelled intermediate PN to PL. In line with this, high PQ supply

favored end product formation in relation to intermediate formation (Fig. 4-28

C). Thus, similarities with PDS are significant in terms of kinetic properties.

5.2.2 OsZDS-His6 might not employ homooligomerization and substrate

channeling

A difference between OsZDS-His6 and OsPDS-His6 came to light upon GPC

analysis. OsZDS-His6 remained flavinylated and is most likely active as

monomer (see 4.4.2.3) whereas OsPDS-His6 is only flavinylated and active as

Page 154: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

140

homotetramer (see 4.2.2 and 5.1.5). It remains to be investigated whether

OsZDS-His6 forms homooligomers at membrane surfaces to employ substrate

channeling of its intermediate PN. In analogy to OsPDS-His6, the size and

oligomeric state of OsZDS-His6 in its active state on liposomal surfaces could

be determined by cryo scanning electron microscopy. Notably, certain

amounts of higher-order oligomers observed during GPC only under non-

reducing conditions might point towards unstable homooligomerization (see

4.4.2.3). Breitenbach et al. (1999) observed homodimerization for Capsicum

ZDS, stating that this might be due to hydrophobic aggregation.

Mathematical modeling of OsZDS-His6 reaction time courses could be applied

to further investigate oligomerization and substrate channeling. Given the

strong similarity of PDS and ZDS, the “monomeric model” and “substrate

channeling model” for OsPDS-His6 might be applicable to OsZDS-His6.

Should OsZDS-His6 not assemble as homooligomer, the question arises whether

ZDS is rather a constituent of a heterooligomeric metabolon, e.g. with CRTISO

that is located downstream of ZDS, as suggested by Lundqvist et al. (2017).

Native expression of OsZDS and LeCRTISO (Yu et al., 2011) in E. coli has

already been achieved and bimolecular fluorescence complementation

(BiFC) assays could help revealing structural interactions. Alternatively, a

different organization might facilitate intermediate release from ZDS to serve

other functions. It has been proposed that a cis-neurosporene might serve as

precursor for a signaling compound involved in the feedback regulation of

carotenogenesis (Kachanovsky et al., 2012). In fact, as will be discussed in

5.4, 9-cis-neurosporene that is formed by CRTISO from proneurosporene (Yu et

al., 2011) can be converted by AtCCD7 to yield 9-cis-ζ-apo-10’-carotenal as

a potential signal precursor.

Page 155: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

141

5.3 ζ-Carotene Isomerase ZISO: a bona fide enzyme with the

potential for additional functions

In the framework of this thesis, functional expression in E. coli and a liposomal

in vitro assay were established for OsZISO-His6 (see 4.5.1). N-terminal fusion of

the Mistic protein from Bacillus subtilis, a fusion partner mediating membrane

targeting and insertion in E. coli (Roosild et al., 2005), was crucial for obtaining

active OsZISO-His6. This is in accordance with ZISO being a transmembrane

protein, as predicted (Chen et al., 2010). Cell lysates containing Mistic-OsZISO-

His6 isomerized 9,15,9‘-tri-cis-ζ-carotene to 9,9‘-di-cis-ζ-carotene in the

absence of all other carotenogenic enzymes (Fig. 4-31) showing that ZISO is

indeed an enzyme rather than a modifier of PDS specificity, the latter being

capable of isomerizing the C9-C10 double bond during phytoene

desaturation (Fig. 1-3 A). Enzyme function was then shown by Beltrán et al.

(2015) who reported its first native purification and in vitro isomerization of tri-

cis-ζ-carotene shortly afterwards. Beltran et al. (2015) furthermore reported on

the cofactor requirements and mechanistics of ZISO, involving a ferrous heme

b cofactor. Their findings are not further detailed here and prompted a

discontinuation of OsZISO-His6 purification attempts during this thesis.

Nevertheless, the data presented here allow first insights into substrate

recognition by ZISO. Apart from 9,15,9‘-tri-cis-ζ-carotene, 9,15-di-cis-

phytofluene was shown to be readily isomerized by Mistic-OsZISO-His6, in

contrast to 15-cis-phytoene (see 4.5.1). It can be concluded that a petaene

and a 9,15-di-cis configuration are the minimal substrate requirements. This

would correspond to recognition of only one half side of the symmetric,

canonical substrate 9,15,9’-tri-cis-ζ-carotene. But further experimentation is

needed. Carotene isomerization can be reversible as shown for the cis-trans-

ß-carotene isomerase D27 (Bruno and Al-Babili, 2016). Reversibility from 9,9’-di-

cis- to 9,15,9’-tri-cis-ζ-carotene would inter alia imply that 9-cis configuration

suffices for recognition by ZISO. Accordingly, 9-cis-isomers of other linear

carotenes (ζ-carotene, neurosporene and lycopene) could be potential ZISO

Page 156: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

142

substrates as well, depending on the contribution of the polyene system to

ZISO substrate specificity. These questions surely need to be addressed in vitro.

Future research should focus on ZISO topology and protein-protein

interactions in the light of suggestions in favor of a carotenogenic

supercomplex achieving metabolite channeling (Cunningham and Gantt,

1998; Shumskaya et al., 2013; Nisar et al., 2015). ZISO is the only

transmembrane protein of the poly-cis carotene desaturation pathway while

all other enzymes interact monotopically with membranes (Gemmecker et

al., 2015; Albrecht et al., 1995; Yu et al., 2011; Yu and Beyer, 2012). ZISO would

be a predestined adaptor to organize the metabolon by protein-protein

interaction. Such a role was for instance found for the transmembrane protein

Erg28P in sterol biosynthesis (Mo and Bard, 2005). First observations may point

towards such a role: Lundqvist et al. (2017) used a BN-PAGE – mass

spectrometry approach and reported that ZISO and PDS as well as CRTISO

and ZDS might be arranged in two heterocomplexes in Arabidopsis.

Additionally, the data presented point towards substrate channeling

between PDS and ZISO in vitro (see 4.5.3). In the absence of OsPDS-His6, when

9,15,9’-tri-cis-ζ-carotene was deposited in liposomes as substrate, Mistic-

OsZISO-His6 activity was low. In contrast, the identical amount of enzyme

mediated much more isomerization (on a relative and absolute scale) when

the substrate was simultaneously produced by OsPDS-His6. This suggests that

ZISO gains more readily access to its substrate in the presence of OsPDS-His6,

conceivably due to substrate channeling. Mass spectrometry approaches

represent a viable option for the identification and characterization of the

suggested membrane-bound carotenogenic metabolon, as demonstrated

for the peroxisomal importomer (Oeljeklaus et al., 2012).

Page 157: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

143

5.4 The poly-cis pathway of carotene desaturation and retrograde

signaling

5.4.1 AtCCD7 forms linear 9-cis-apocarotenoids as potential signaling

molecule precursors

In recent years, apocarotenoids originating from CCD-mediated cleavage of

poly-cis-configured linear carotenes have been suggested to serve regulatory

functions in plants (see 1.4). Such ideas are of special interest in the light of a

significant conundrum in carotenoid research: why has the highly intricate

poly-cis pathway of carotene desaturation evolved in plants although its

overall reaction can be achieved with only one enzyme, as in bacteria and

fungi (Fig. 1-2)? Sensing desaturation intermediate levels and translating these

into regulatory functions could provide a clue. These lines of thinking have

been fostered by the discovery that strigolactones and abscisic acid, both

important phytohormones, are derived from the cleavage of cis-configured

carotenoids. Avendaño-Vázquez et al. (2014) postulated cis-ζ-carotene or cis-

phytofluene to be cleaved by CCD4 to yield an apocarotenoid participating

in plastid retrograde signaling that regulates seedling development in

Arabidopsis. Kachanovsky et al. (2012) suggested prolycopene,

proneurosporene or a cis-neurosporene to give rise to an apocarotenoid

mediating feedback regulation of fruit-specific PSY1 in tomato.

The lack of evidence for the existence of such linear apocarotenoids,

prompted the here presented research on CCD-mediated cleavage of

canonical intermediates of the poly-cis pathway of carotene desaturation

and non-canonical isomers derived thereof (see 4.6). The carotenoid

cleavage dioxygenases AtCCD4 and AtCCD7 were investigated for primary

cleavage since they are plastid localized (the site of carotene desaturation)

and they are known to cleave cyclic C40 carotenes (Bruno et al., 2016; Bruno

et al., 2014). In contrast, CCD8 does not cleave C40 carotenoids (Alder et al.,

2012) and CCD1 is localized in the cytosol (Floss and Walter, 2009).

Page 158: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

144

Contrary to expectation (Avendaño-Vázquez et al., 2014), AtCCD4, did not

cleave of any cis- or trans-configured linear carotene (see 4.6.1). This is in

accordance with the reported specificity of AtCCD4 for trans-configured

cyclic carotenoids (Bruno et al., 2016). Thus, AtCCD4 seems to require the

cyclic ionone moieties for substrate recognition. The data presented in this

work rule out that AtCCD4 cleaves cis-ζ-carotenes. The only possibility for

AtCCD4 involvement would therefore relate to a secondary cleavage of an

apocarotenoid originating from a desaturation intermediate.

In contrast, AtCCD7 was found to specifically cleave the trans-configured C9-

C10 double bond in 9-cis-ζ-carotene, 9’-cis-neurosporene and 9-cis-lycopene

whereas poly-cis- and all-trans-configured species were not converted (see

4.6.2). This is in accordance with the 9-cis-specificity of AtCCD7 for bicyclic

carotenoids (Bruno et al., 2014) that appears to persist with acyclic carotenes.

The apocarotenoids formed by AtCCD7 were identified as 9-cis-ζ-apo-10’-

carotenal (from ζ-carotene and neurosporene) and 9-cis-apo-10’-lycopenal

(from lycopene) by LC-MS and the stereo-configuration of the lycopenal was

verified by its chromatographic behavior differing from authentic all-trans-

apo-10’-lycopenal. Cleavage of the 9-cis isomers of neurosporene and

lycopene was comparably weaker than cleavage of 9-cis-ζ-carotene. This is

somewhat surprising considering the fact that they resemble the canonical

CCD7 substrate 9-cis-β-carotene much more with respect to the number of

double bonds present. The explanation might be that chromophore

elongation hinders single bond rotation within the polyene. The resulting

enhanced rigidity is known to cause problems upon solubilization with

detergents so that more desaturated substrates like neurosporene and

lycopene are not optimally presented to the CCD enzymes.

The findings of this work are reminiscent of SL biosynthesis where CCD7

cleaves 9-cis-β-carotene at the trans-configured C9-C10 double bond to

yield a 9-cis-C27-apocarotenoid (9-cis-β-apo-10’-carotenal) (Al-Babili and

Bouwmeester, 2015). Moreover, ABA as the second carotenoid-derived

Page 159: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

145

phytohormone has a 9-cis-carotenoid precursor, too (Marion-Poll and

Nambara, 2005). There seems to be the recurrent motif of a 9-cis-configured

carotenoid representing precursors of regulatory molecules. Following this

idea, the data presented in this work support an involvement of AtCCD7 in

the biosynthesis of regulatory molecule derived directly or indirectly from the

poly-cis pathway of carotene desaturation.

Interestingly, 9-cis configuration is widespread amongst carotenes of the poly-

cis pathway: 9-cis double bonds are introduced at both C9-C10 and C9’-C10’

by PDS (Fig. 1-3 A) and persist throughout the desaturation pathway

intermediates until being isomerized to trans by CRTISO (see Fig. 1-2). Thus, all

of the poly-cis pathway intermediates might be predestined as precursors of

regulatory molecules. Moreover, 9-mono-cis-isomers of linear carotenes might

indeed be formed in planta. 9’-cis-neurosporene and 9-cis-lycopene can be

enzymatically formed in vitro by the carotene isomerase CRTISO (Isaacson et

al., 2004; Yu et al., 2011), a property that has been exploited in this work to

generate 9’-cis-neurosporene as substrate. In contrast, the potential origin of

9-cis-ζ-carotene in planta remains elusive. Taking the resemblance with SL

biosynthesis into account, the canonical 9,9’-di-cis-ζ-carotene should be

assayed in vitro as substrate with the 9-cis-β-carotene isomerase D27. Further

candidates could be the yet uncharacterized CRTISO-like enzymes (Fantini et

al., 2013) that might represent carotene isomerases with the required

specificity. Moreover, thermo- and photoisomerization should be considered.

9-cis-ζ-carotene and other isomers were regularly found in E. coli (see 4.5.1), in

the absence of specific carotene isomerases.

5.4.2 Linear 9-cis-apocarotenoids are not converted into strigolactone-like

metabolites by AtCCD8

Signaling molecules need to move between aqueous cell compartments in

order to exert their functions, a feature that does not apply to the rather

hydrophobic C27 apocarotenoids 9-cis-ζ-apo-10’-carotenal and 9-cis-apo-10’-

lycopenal. Further modification or cleavage by CCDs should be considered.

Page 160: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Discussion

146

Given a potential resemblance to SL biosynthesis, CCD8 would be the prime

candidate to potentially yield strigolactone-like compounds from the linear

apocarotenoids (Alder et al., 2012). However, no conversion was observed

(Fig. 4-36) which is somewhat surprising: The postulated CCD8 reaction

mechanism for carlactone formation from 9-cis-β-apo-10’-carotenal (Alder et

al., 2012) involves only C11=C12 and C13=C14 to participate directly in the

intramolecular rearrangement being surrounded by C15=C15’ and C9=C10.

All of these double bonds as well as 9-cis-configuration are present in 9-cis-ζ-

apo-10’-carotenal and 9-cis-apo-10’-lycopenal, the latter being equipped

with a polyene that strongly resembles β-apo-10-carotenal. It remains to be

assayed whether the lycopenal can be converted by AtCCD8. However, the

absence of the ionone moiety might hinder recognition by AtCCD8.

Consequently, the newly identified apocarotenoids 9-cis-ζ-apo-10’-carotenal

and 9-cis-apo-10’-lycopenal might rather undergo modification by hitherto

unknown enzyme activities. Accordingly, conversion of 9-cis-ζ-apo-10’-

carotenal and 9-cis-apo-10’-lycopenal by AtCCD1 and AtCCD4, both known

to convert apocarotenoids (Bruno et al., 2016; Floss and Walter, 2009) should

be considered. However, AtCCD4 did not cleave the ζ-carotenal and

cleavage by AtCCD1 did not yield a cleavage product that was visible in the

UV/VIS wavelength range from 250 nm to 400 nm (Fig. 4-36). Considering the

broad cleavage site specificity of CCD1 enzymes yielding diverse products

from one substrate (Floss and Walter, 2009), it seems unlikely that a CCD1-

mediated reaction might yield a regulatory molecule.

The history of low molecular mass signaling molecule identification shows that

plant mutant characterization might help to further elucidate the fate of

these two apocarotenoids. Alternatively, the classic approach of synthesizing

radioactivity-labeled apocarotenoids, track their fates upon application to

plant cells and identify their derivatives by LC-MS and NMR analyses remains

a viable option.

Page 161: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Summary

147

6 Summary

Plant carotenoid biosynthesis in plastid membranes proceeds via the so-

called poly-cis pathway of carotene desaturation, converting 15-cis-

phytoene into all-trans-lycopene via poly-cis-configured intermediates. Its four

constituent enzymes are poorly understood in terms of mechanisms and

structure. In addition, the reason for the prevalence of these intricate

reactions remains enigmatic considering that the bacterial carotene

desaturation pathway requires only one enzyme.

In this work, the crystal structure of phytoene desaturase (PDS) from O. sativa,

the first desaturation pathway enzyme, in complex with its inhibitor norflurazon

(NFZ) was refined and structural implications were functionally evaluated by a

kinetic characterization. The kinetics of the downstream ζ-carotene

desaturase (ZDS), representing a PDS homolog, were investigated for

comparison and revealed pronounced similarities. The data presented

support an ordered ping pong bi bi kinetic mechanism for PDS and ZDS.

Desaturation of their substrates occurs via two formally identical,

mechanistically independent reactions on half sides of the symmetric

substrate. The intermediate is released into the membrane after the first

desaturation. The enzyme-bound FADred produced by carotene desaturation

is reoxidized by plastoquinone (PQ) to allow repeated rounds of catalysis. As

inferred from the active site structure, data support an unprecedented “flavin

only” mechanism characterized by FADox as sole catalyst of carotene

desaturation. PDS and ZDS interact monotopically with membranes to access

their substrates. Mathematical modelling implies substrate channeling

between subunits within PDS homotetramers whereas ZDS is active as

monomer. Notably, PDS does not produce ζ-carotene in the

stereoconfiguration required by ZDS. In this work, a cell-free activity assay

demonstrated that the membrane-integral ζ-carotene isomerase (ZISO) is a

bona fide isomerase representing the link between PDS and ZDS.

Page 162: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Summary

148

An additional aspect of this work relates to the mode of action of bleaching

herbicides targeting PDS and containing meta-CF3-phenyl moieties, such as

NFZ. It is demonstrated that they compete with PQ and NFZ, occupying the

PQ binding site as suggested in crystallo. The meta-CF3 substituent provides

unique non-covalent interactions with conserved residues and is crucial for

inhibitory effectiveness. Mutagenesis of the conserved Arg300, forming a

hydrogen bond with NFZ, allows engineering NFZ-resistant PDS but at the

expense of altered enzyme kinetics and partial loss of enzymatic activity.

Evidence has been presented that carotenoid cleavage dioxygenases

(CCDs) might form retrograde signals from cis-configured carotene

desaturation intermediates, regulating plastid development and carotenoid

biosynthesis. Investigating intermediate cleavage by CCDs, Arabidopsis CCD7

was found to cleave 9-cis-ζ-carotene yielding 9-cis-ζ-apo-10’-carotenal (C27)

which, in contrast to strigolactone biosynthesis, is not converted by AtCCD8.

These data suggest this apocarotenoid as a candidate signal precursor

requiring further modification by hitherto unidentified enzyme activities.

Page 163: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

References

149

7 References

Al-Babili S, Hugueney P, Schledz M, Welsch R, Frohnmeyer H, Laule O, Beyer P (2000)

Identification of a novel gene coding for neoxanthin synthase from Solanum tuberosum. FEBS

Lett 485: 168–72

Al-Babili S, von Lintig J, Haubruck H, Beyer P (1996) A novel, soluble form of phytoene

desaturase from Narcissus pseudonarcissus chromoplasts is Hsp70-complexed and

competent for flavinylation, membrane association and enzymatic activation. Plant J 9: 601–

612

Albrecht M, Klein A, Hugueney P, Sandmann G, Kuntz M (1995) Molecular cloning and

functional expression in E. coli of a novel plant enzyme mediating ζ-carotene desaturation.

FEBS Lett 372: 199–202

Alder A, Jamil M, Marzorati M, Bruno M, Vermathen M, Bigler P, Ghisla S, Bouwmeester H, Beyer P, Al-Babili S (2012) The Path from β-carotene to Carlactone, a Strigolactone-Like Plant

Hormone. Science 335: 1348–1351

Álvarez D, Voß B, Maass D, Wüst F, Schaub P, Beyer P, Welsch R (2016) Carotenogenesis Is

Regulated by 5′UTR-Mediated Translation of Phytoene Synthase Splice Variants. Plant Physiol.

172: 2314-2326

Arango J, Jourdan M, Geoffriau E, Beyer P, Welsch R (2014) Carotene Hydroxylase Activity

Determines the Levels of Both β-Carotene and Total Carotenoids in Orange Carrots. Plant Cell

26: 2223–2233

Arias RS, Dayan FE, Michel A, Howell J, Scheffler BE (2006) Characterization of a higher plant

herbicide-resistant phytoene desaturase and its use as a selectable marker. Plant Biotechnol

J 4: 263–273

Arnoux P, Morosinotto T, Saga G, Bassi R, Pignol D (2009) A structural basis for the pH-

dependent xanthophyll cycle in Arabidopsis thaliana. Plant Cell 21: 2036–2044

Avendaño-Vázquez A-O, Cordoba E, Llamas E, San Román C, Nisar N, De la Torre S, Ramos-Vega M, Gutiérrez-Nava MDLL, Cazzonelli CI, Pogson BJ, León P (2014) An Uncharacterized

Apocarotenoid-Derived Signal Generated in ζ-Carotene Desaturase Mutants Regulates Leaf

Development and the Expression of Chloroplast and Nuclear Genes in Arabidopsis. Plant Cell

26: 2524–2537

Babczinski P, Sandmann G, Schmidt RR, Shiokawa K, Yasui K (1995) Substituted

Tetrahydropyrimidinones: A New Herbicidal Class of Compounds Inducing Chlorosis by

Inhibition of Phytoene Desaturation. Pestic Biochem Physiol 52: 33–44

Bai L, Kim E-H, DellaPenna D, Brutnell TP (2009) Novel lycopene ε-cyclase activities in maize

revealed through perturbation of carotenoid biosynthesis. Plant J 59: 588–599

Bartley GE, Scolnik PA, Beyer P (1999) Two Arabidopsis thaliana carotene desaturases,

phytoene desaturase and ζ-carotene desaturase, expressed in Escherichia coli, catalyze a

poly-cis pathway to yield pro-lycopene. Eur J Biochem 259: 396–403

Bartnikas TB, Tosques IE, Laratta WP, Shi J, Shapleigh JP (1997) Characterization of the nitric

oxide reductase-encoding region in Rhodobacter sphaeroides 2.4.3. J Bacteriol 179: 3534–

3540

Beltrán J, Kloss B, Hosler JP, Geng J, Liu A, Modi A, Dawson JH, Sono M, Shumskaya M, Ampomah-Dwamena C, Love JD, Wurtzel ET (2015) Control of carotenoid biosynthesis through

a heme-based cis-trans isomerase. Nat Chem Biol 11: 598–605

Beyer P, Al-Babili S, Ye X, Lucca P, Schaub P, Welsch R, Potrykus I (2002) Golden Rice:

introducing the β-carotene biosynthesis pathway into rice endosperm by genetic engineering

to defeat vitamin A deficiency. J Nutr 132: 506S–510S

Page 164: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

References

150

Beyer P, Mayer M, Kleinig H (1989) Molecular oxygen and the state of geometric isomerism of

intermediates are essential in the carotene desaturation and cyclization reactions in daffodil

chromoplasts. Eur J Biochem 184: 141–150

Bonk M, Hoffmann B, Lintig J, Schledz M, Al-Babili S, Hobeika E, Kleinig H, Beyer P (1997)

Chloroplast Import of Four Carotenoid Biosynthetic Enzymes In Vitro Reveals Differential Fates

Prior to Membrane Binding and Oligomeric Assembly. Eur J Biochem 247: 942–950

Booker J, Auldridge M, Wills S, McCarty D, Klee H, Leyser O (2004) MAX3/CCD7 Is a

Carotenoid Cleavage Dioxygenase Required for the Synthesis of a Novel Plant Signaling

Molecule. Curr Biol 14: 1232–1238

Bouvier F, D’Harlingue A, Backhaus RA, Kumagai MH, Camara B (2000) Identification of

neoxanthin synthase as a carotenoid cyclase paralog. Eur J Biochem 267: 6346–6352

Bouvier F, D’Harlingue A, Hugueney P, Marin E, Marion-Poll A, Camara B (1996) Xanthophyll

Biosynthesis. J Biol Chem 271: 28861–28867

Bramley PM (2002) Regulation of carotenoid formation during tomato fruit ripening and

development. J Exp Bot 53: 2107–2113

Brausemann A, Gemmecker S, Koschmieder J, Ghisla S, Beyer P, Einsle O (2017) Structure of

Rice (O. sativa) Phytoene Desaturase Provides Insights Into Herbicide Binding and Reaction

Mechanisms Involved In Carotene Desaturation, Structure, in press

Breitenbach J, Bruns M, Sandmann G (2013) Contribution of two ζ-carotene desaturases to

the poly-cis desaturation pathway in the cyanobacterium Nostoc PCC 7120. Arch Microbiol

195: 491–498

Breitenbach J, Fernández-González B, Vioque A, Sandmann G (1998) A higher-plant type ζ-

carotene desaturase in the cyanobacterium Synechocystis PCC6803. Plant Mol Biol 36: 725–

732

Breitenbach J, Kuntz M, Takaichi S, Sandmann G (1999) Catalytic properties of an expressed

and purified higher plant type ζ-carotene desaturase from Capsicum annuum. Eur J Biochem

265: 376–383

Breitenbach J, Sandmann G (2005) ζ-Carotene cis isomers as products and substrates in the

plant poly-cis carotenoid biosynthetic pathway to lycopene. Planta 220: 785–793

Breitenbach J, Zhu C, Sandmann G (2001) Bleaching herbicide norflurazon inhibits phytoene

desaturase by competition with the cofactors. J Agric Food Chem 49: 5270–5272

Britton G, Liaaen-Jensen S, Pfander H (2004) Carotenoids - Handbook. 1st ed., Birkhäuser, Basel

Britton G, Liaaen-Jensen S, Pfander H (2004) Carotenoids - Volume 1 B: Spectroscopy, 1st ed.,

Birkhäuser, Basel

Bruno M, Al-Babili S (2016) On the substrate specificity of the rice strigolactone biosynthesis

enzyme DWARF27. Planta 243: 1429–1440

Bruno M, Hofmann M, Vermathen M, Alder A, Beyer P, Al-Babili S (2014) On the substrate- and

stereospecificity of the plant carotenoid cleavage dioxygenase 7. FEBS Lett 588: 1802–1807

Bruno M, Koschmieder J, Wüst F, Schaub P, Fehling-Kaschek M, Timmer J, Beyer P, Al-Babili S

(2016) Enzymatic study on AtCCD4 and AtCCD7 and their potential to form acyclic

regulatory metabolites. J Exp Bot 67: 5993–6005

Bruno M, Vermathen M, Alder A, Wüst F, Schaub P, van der Steen R, Beyer P, Ghisla S, Al-Babili S (2017) Insights into the formation of carlactone from in-depth analysis of the CCD8-

catalyzed reactions. FEBS Lett 591: 792–800

Bugos RC, Hieber AD, Yamamoto HY (1998) Xanthophyll cycle enzymes are members of the

lipocalin family, the first identified from plants. J Biol Chem 273: 15321–15324

Carol P, Kuntz M (2001) A plastid terminal oxidase comes to light: Implications for carotenoid

Page 165: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

References

151

biosynthesis and chlororespiration. Trends Plant Sci 6: 31–36

Cazzonelli CI (2011) Carotenoids in nature: Insights from plants and beyond. Funct Plant Biol

38: 833–847

Cazzonelli CI, Pogson BJ (2010) Source to sink: regulation of carotenoid biosynthesis in plants.

Trends Plant Sci 15: 266–274

Chamovitz D, Sandmann G, Hirschberg J (1993) Molecular and biochemical characterization

of herbicide-resistant mutants of cyanobacteria reveals that phytoene desaturation is a rate-

limiting step in carotenoid biosynthesis. J Biol Chem 268: 17348–17353

Chen Y, Li F, Wurtzel ET (2010) Isolation and characterization of the Z-ISO gene encoding a

missing component of carotenoid biosynthesis in plants. Plant Physiol 153: 66–79

Clough JM, Pattenden G (1979) Naturally occurring poly-cis carotenoids. Stereochemistry of

poly-cis lycopene and its congeners in “Tangerine” tomato fruits. J Chem Soc, Chem

Commun 616–619

Copeland RA (2000), Enzymes: a practical introduction to structure, mechanism and data

analysis, 1st ed., Wiley-VCH, Weinheim

Cramp MC, Gilmour J, Hatton LR, Hewett RH, Nolan CJ, Parnell EW (1987) Design and Synthesis

of N-(2,4-difluorophenyl)-2-(3-trifluoromethylphenoxy)-3-pyridinecarboxamide (diflufenican), a

Novel Pre- and Early Post-emergence Herbicide for Use in Winter Cereals. Pestic Sci 18: 15–28

Cunningham FX, Gantt E (1998) Genes and Enzymes of Carotenoid Biosynthesis in Plants. Annu

Rev Plant Physiol Plant Mol Biol 49: 557–583

Cunningham FX, Pogson B, Sun Z, McDonald K a, DellaPenna D, Gantt E (1996) Functional

Analysis of the β and ε Lycopene Cyclase Enzymes of Arabidopsis Reveals a Mechanism for

Control of Cyclic Carotenoid Formation. Plant Cell 8: 1613–1626

Cuttriss AJ, Chubb AC, Alawady A, Grimm B, Pogson BJ (2007) Regulation of lutein

biosynthesis and prolamellar body formation in Arabidopsis. Funct Plant Biol 34: 663–672

Degli Esposti M, Bertoli E, Parenti-Castelli G, Fato R, Mascarello S, Lenaz G (1981) Incorporation

of Ubiquinone Homologs into Lipid Vesicles and Mitochondrial Membranes. Arch Biochem

Biophys 210: 21–32

Demmig-Adams B, Gilmore AM, Adams WW (1996) In vivo functions of carotenoids in higher

plants. Faseb 10: 403–412

Diretto G, Tavazza R, Welsch R, Pizzichini D, Mourgues F, Papacchioli V, Beyer P, Giuliano G

(2006) Metabolic engineering of potato tuber carotenoids through tuber-specific silencing of

lycopene ε-cyclase. BMC Plant Biol 6: 13

Dogbo O, Laferriere A, d’Harlingue A, Camara B (1988) Carotenoid Biosynthesis: Isolation and

Characterization of a Bifunctional Enzyme Catalyzing the Synthesis of Phytoene. Proc Natl

Acad Sci USA 85: 7054–7058

Dong H, Deng Y, Mu J, Lu Q, Wang Y, Xu Y, Chu C, Chong K, Lu C, Zuo J (2007) The Arabidopsis

Spontaneous Cell Death1 gene, encoding a ζ-carotene desaturase essential for carotenoid

biosynthesis, is involved in chloroplast development, photoprotection and retrograde

signalling. Cell Res 17: 458–470

Ernst S, Sandmann G (1988) Poly-cis carotene pathway in the Scenedesmus mutant C-6D.

Arch Microbiol 150: 590–594

Fantini E, Falcone G, Frusciante S, Giliberto L, Giuliano G (2013) Dissection of Tomato

Lycopene Biosynthesis through Virus-Induced Gene Silencing. Plant Physiol 163: 986–998

Floss DS, Walter MH (2009) Role of carotenoid cleavage dioxygenase 1 (CCD1) in

apocarotenoid biogenesis revisited. Plant Signal Behav 4: 172–175

Fraser PD, Misawa N, Linden H, Yamano S, Kobayashi K, Sandmann G (1992) Expression in

Page 166: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

References

152

Escherichia coli, purification, and reactivation of the recombinant Erwinia uredovora

phytoene desaturase. J Biol Chem 267: 19891–19895

Fufezan C, Simionato D, Morosinotto T (2012) Identification of key residues for pH dependent

activation of violaxanthin de-epoxidase from Arabidopsis thaliana. PLoS One 7: e35669

Gemmecker S, Schaub P, Koschmieder J, Brausemann A, Drepper F, Rodriguez-Franco M, Ghisla S, Warscheid B, Einsle O, Beyer P (2015) Phytoene Desaturase from Oryza sativa:

Oligomeric Assembly, Membrane Association and Preliminary 3D-Analysis. PLoS One 10:

e0131717

Gemmecker S (2015) Biochemische und strukturelle Charaktierisierung der Phytoen-

Desaturasen aus Pantoea ananatis und Oryza sativa. University of Freiburg, Germany

Gibson DG, Young L, Chuang R-Y, Venter JC, Hutchison CA, Smith HO (2009) Enzymatic

assembly of DNA molecules up to several hundred kilobases. Nat Methods 6: 343–345

Giuliano G, Bartley GE, Scolnik PA (1993) Regulation of Carotenoid Biosynthesis during Tomato

Development. Plant Cell 5: 379–387

Greenwood AI, Tristram-Nagle S, Nagle JF (2006) Partial molecular volumes of lipids and

cholesterol. Chem Phys Lipids 143: 1–10

Gruszecki WI, Strzalka K (2005) Carotenoids as modulators of lipid membrane physical

properties. Biochim Biophys Acta - Mol Basis Dis 1740: 108–115

Hagemann GE, Sturgis JN, Wagner E, Robert B, Beyer P, Tadros MH (1996) The effects of the

detergent LDAO on the carotenoid metabolism and growth of Rhodovulum sulfidophilum.

Microbiol Res 151: 421–426

Hanukoglu I (2015) Rossmann fold: A β-α-βfold at dinucleotide binding sites. Biochem Mol Biol

Educ 43: 206–209

Havaux M (1998) Carotenoids as membrane stabilizers in chloroplasts. Trends Plant Sci 3: 147–

151

Havaux M (2014) Carotenoid oxidation products as stress signals in plants. Plant J 79: 597–606

Heckman KL, Pease LR (2007) Gene splicing and mutagenesis by PCR-driven overlap

extension. Nat Protoc 2: 924–932

Hengartner U, Bernhard K, Meyer K, Englert G, Glinz E (1992) Synthesis, Isolation, and NMR-

Spectroscopic Characterization of Fourteen (Z)-Isomers of Lycopene and of Some Acetylenic

Didehydro- and Tetradehydrolycopenes. Helv Chim Acta 75: 1848–1865

Hou X, Rivers J, León P, McQuinn RP, Pogson BJ (2016) Synthesis and Function of

Apocarotenoid Signals in Plants. Trends Plant Sci 21: 792–803

Huang FC, Molnár P, Schwab W (2009) Cloning and functional characterization of carotenoid

cleavage dioxygenase 4 genes. J Exp Bot 60: 3011–3022

Huang J, Liu J, Li Y, Chen F (2008) Isolation and characterization of the phytoene desaturase

gene as a potential selective marker for genetic engineering of the astaxanthin-producing

green alga Chlorella zofingiensis (Chlorophyta). J Phycol 44: 684–690

Hugueney P, Römer S, Kuntz M, Camara B (1992) Characterization and molecular cloning of a

flavoprotein catalyzing the synthesis of phytofluene and ζ-carotene in Capsicum

chromoplasts. Eur J Biochem 209: 399–407

Isaacson T, Ohad I, Beyer P, Hirschberg J (2004) Analysis in vitro of the enzyme CRTISO

establishes a poly-cis-carotenoid biosynthesis pathway in plants. Plant Physiol 136: 4246–4255

Isaacson T, Ronen G, Zamir D, Hirschberg J (2002) Cloning of tangerine from tomato reveals a

carotenoid isomerase essential for the production of β-carotene and xanthophylls in plants.

Plant Cell 14: 333–342

Page 167: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

References

153

Johnson ER, Keinan S, Mori-Sánchez P, Contreras-García J, Cohen AJ, Yang W (2010)

Revealing noncovalent interactions. J Am Chem Soc 132: 6498–6506

Josse EM, Alcaraz JP, Labouré AM, Kuntz M (2003) In vitro characterization of a plastid

terminal oxidase (PTOX). Eur J Biochem 270: 3787–3794

Joyard J, Ferro M, Masselon C, Seigneurin-Berny D, Salvi D, Garin J, Rolland N (2009)

Chloroplast proteomics and the compartmentation of plastidial isoprenoid biosynthetic

pathways. Mol Plant 2: 1154–1180

Kachanovsky DE, Filler S, Isaacson T, Hirschberg J (2012) Epistasis in tomato color mutations

involves regulation of phytoene synthase 1 expression by cis-carotenoids. Proc Natl Acad Sci

109: 19021–19026

Kaschek D, Mader W, Fehling-Kaschek M, Rosenblatt M, Timmer J (2016) Dynamic Modeling,

Parameter Estimation and Uncertainty Analysis in R. bioRxiv

Kim J, Smith JJ, Tian L, DellaPenna D (2009) The evolution and function of carotenoid

hydroxylases in Arabidopsis. Plant Cell Physiol 50: 463–479

Kowalczyk-Schröder, Susanne; Sandmann G (1992) Interference of Fluridone with the

Desaturation of Phytoene by Membranes of the Cyanobacterium Aphanocapsa. Pestic

Biochem Physiol 42: 7–12

Laber B, Usunow G, Wiecko E, Franke W, Franke H (1999) Inhibition of Narcissus

pseudonarcissus Phytoene Desaturase by Herbicidal 3-Trifluoromethyl-1,1-biphenyl Derivatives.

184: 173–184

Lee H, Deloache WC, Dueber JE (2012) Spatial organization of enzymes for metabolic

engineering. Metab Eng 14: 242–251

Li F, Murillo C, Wurtzel ET (2007) Maize Y9 encodes a product essential for 15-cis-ζ-carotene

isomerization. Plant Physiol 144: 1181–1189

Linden H (1993) Isolation of a carotenoid biosynthesis gene coding for ζ-carotene desaturase

from Anabaena PCC 7120 by heterologous complementation. FEMS Microbiol Lett 106: 99–

103

Linden H, Misawa N, Chamovitz D, Pecker I, Hirschberg J, Sandmann G (1991) Functional

Complementation in Escherichia coli of Different Phytoene Desaturase Genes and Analysis of

Accumulated Carotenes. Z Naturforsch 46 c: 1045–1051

Liu J, Gerken H, Huang J, Chen F (2013) Engineering of an endogenous phytoene desaturase

gene as a dominant selectable marker for Chlamydomonas reinhardtii transformation and

enhanced biosynthesis of carotenoids. Process Biochem 48: 788–795

Liu J, Zhong Y, Sun Z, Huang J, Sandmann G, Chen F (2010) One amino acid substitution in

phytoene desaturase makes Chlorella zofingiensis resistant to norflurazon and enhances the

biosynthesis of astaxanthin. Planta 232: 61–67

Lopez AB, Yang Y, Thannhauser TW, Li L (2008) Phytoene desaturase is present in a large

protein complex in the plastid membrane. Physiol Plant 133: 190–198

Lundquist PK, Mantegazza O, Stefanski A, Stühler K, Weber APM (2017) Surveying the

Oligomeric State of Arabidopsis thaliana Chloroplasts. Mol Plant 10: 197–211

Martinez-Férez I, Vioque A, Sandmann G (1994) Mutagenesis of an Amino Acid Responsible in

Phytoene Desaturase from Synechocystis for Binding of the Bleaching Herbicide Norflurazon.

Pestic Biochem Physiol 48: 185–190

Mayer MP, Bartlett DL, Beyer P, Kleinig H (1989) The in vitro mode of action of bleaching

herbicides on the desaturation of 15-cis-phytoene and cis-ζ-carotene in isolated daffodil

chromoplasts. Pestic Biochem Physiol 34: 111–117

Mayer MP, Beyer P, Kleinig H (1990) Quinone compounds are able to replace molecular

Page 168: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

References

154

oxygen as terminal electron acceptor in phytoene desaturation in chromoplasts of Narcissus

pseudonarcissus L. Eur J Biochem 191: 359–363

McDonald AE, Ivanov AG, Bode R, Maxwell DP, Rodermel SR, Hüner NPA (2011) Flexibility in

photosynthetic electron transport: The physiological role of plastoquinol terminal oxidase

(PTOX). Biochim Biophys Acta 1807: 954–967

Misawa N, Masamoto K, Hori T, Ohtani T, Böger P, Sandmann G (1994) Expression of an Erwinia

phytoene desaturase gene not only confers multiple resistance to herbicides interfering with

carotenoid biosynthesis but also alters xanthophyll metabolism in transgenic plants. Plant J 6:

481–489

Mo C, Bard M (2005) Erg28p is a key protein in the yeast sterol biosynthetic enzyme complex. J

Lipid Res 46: 1991–1998

Nambara E, Marion-Poll A (2005) Abscisic acid biosynthesis and catabolism. Annu Rev Plant

Biol 56: 165–185

Nievelstein V, Vandekerchove J, Tadros MH, Lintig J V, Nitschke W, Beyer P (1995) Carotene

desaturation is linked to a respiratory redox pathway in Narcissus pseudonarcissus

chromoplast membranes. Involvement of a 23-kDa oxygen-evolving-complex-like protein. Eur

J Biochem 233: 864–872

Nisar N, Li L, Lu S, Khin NC, Pogson BJ (2015) Carotenoid metabolism in plants. Mol Plant 8: 68–

82

Nocedal J, Wright S (1999) Numerical optimization, 1st ed., Springer Science+Business Media,

Berlin

Nogueira M, Mora L, Enfissi EMA, Bramley PM, Fraser PD (2013) Subchromoplast sequestration

of carotenoids affects regulatory mechanisms in tomato lines expressing different carotenoid

gene combinations. Plant Cell 25: 4560–79

Norris SR, Barrette TR, DellaPenna D (1995) Genetic dissection of carotenoid synthesis in

Arabidopsis defines plastoquinone as an essential component of phytoene desaturation.

Plant Cell 7: 2139–2149

North HM, Almeida AD, Boutin JP, Frey A, To A, Botran L, Sotta B, Marion-Poll A (2007) The

Arabidopsis ABA-deficient mutant aba4 demonstrates that the major route for stress-induced

ABA accumulation is via neoxanthin isomers. Plant J 50: 810–824

Ohki S, Miller-Sulger R, Wakabayashi K, Pfleiderer W, Böger P (2003) Phytoene Desaturase

Inhibition by O-(2-Phenoxy)ethyl- N -aralkylcarbamates. J Agric Food Chem 51: 3049–3055

Oeljeklaus S, Reinartz BS, Wolf J, Wiese S, Tonillo J, Podwojski K, Kuhlmann K, Stephan C, Meyer HE, Schliebs W, Brocard C, Erdmann R, Warscheid B (2012) Identification of core

components and transient interactors of the peroxisomal importomer by dual-track stable

isotope labeling with amino acids in cell culture analysis. J Proteome Res 11: 2567–2580

Park H, Kreunen SS, Cuttriss AJ, DellaPenna D, Pogson BJ (2002) Identification of the

carotenoid isomerase provides insight into carotenoid biosynthesis, prolamellar body

formation, and photomorphogenesis. Plant Cell 14: 321–332

Pecker I, Chamovitz D, Linden H, Sandmann G, Hirschberg J (1992) A single polypeptide

catalyzing the conversion of phytoene to ζ-carotene is transcriptionally regulated during

tomato fruit ripening. Proc Natl Acad Sci U S A 89: 4962–4966

Plumley FG, Schmidt GW (1987) Reconstitution of chlorophyll a/b light-harvesting complexes:

Xanthophyll-dependent assembly and energy transfer. Proc Natl Acad Sci U S A 84: 146–150

Qureshi AA, Kim M, Qureshi N, Porter JW (1974) The enzymatic conversion of cis-

(14C)phytofluene, trans-(14C)phytofluene, and trans-ζ-(14C)carotene to poly-cis-acyclic

carotenes by a cell-free preparation of tangerine tomato fruit plastids. Arch Biochem Biophys

162: 108–116

Page 169: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

References

155

Raisig A, Sandmann G (2001) Functional properties of diapophytoene and related

desaturases of C30 and C40 carotenoid biosynthetic pathways. Biochim Biophys Acta - Mol

Cell Biol Lipids 1533: 164–170

Ramel F, Birtic S, Ginies C, Soubigou-Taconnat L, Triantaphylides C, Havaux M (2012)

Carotenoid oxidation products are stress signals that mediate gene responses to singlet

oxygen in plants. Proc Natl Acad Sci 109: 5535–5540

Raue A, Kreutz C, Maiwald T, Bachmann J, Schilling M, Klingmüller U, Timmer J (2009)

Structural and practical identifiability analysis of partially observed dynamical models by

exploiting the profile likelihood. Bioinformatics 25: 1923–1929

Rodriguez-Concepción M, Boronat A (2002) Elucidation of the Methylerythritol Phosphate

Pathway for Isoprenoid Biosynthesis in Bacteria and Plastids. A Metabolic Milestone Achieved

through Genomics. Plant Physiol 130: 1079–1089

Römer S, Fraser PD, Kiano JW, Shipton CA, Misawa N, Schuch W, Bramley PM (2000) Elevation

of the provitamin A content of transgenic tomato plants. Nat Biotechnol 18: 666–669

Römer S, Lübeck J, Kauder F, Steiger S, Adomat C, Sandmann G (2002) Genetic engineering

of a zeaxanthin-rich potato by antisense inactivation and co-suppression of carotenoid

epoxidation. Metab Eng 4: 263–272

Ronen G, Carmel-Goren L, Zamir D, Hirschberg J (2000) An alternative pathway to β-carotene

formation in plant chromoplasts discovered by map-based cloning of Beta and old-gold

color mutations in tomato. Proc Natl Acad Sci 97: 11102–11107

Roosild TP (2005) NMR Structure of Mistic, a Membrane-Integrating Protein for Membrane

Protein Expression. Science (80- ) 307: 1317–1321

Ruyter-Spira C, Al-Babili S, van der Krol S, Bouwmeester H (2013) The biology of strigolactones.

Trends Plant Sci 18: 72–83

Sandmann G (2009) Evolution of carotene desaturation: The complication of a simple

pathway. Arch Biochem Biophys 483: 169–174

Sandmann G, Böger P (1982) Bleaching Activity of New 2-Phenylpyridazinones : Structure-

Activity Relationship. Z Naturforsch 37 c: 1092- 1094

Sandmann G, Böger, P (1997) Phytoene desaturase as target for bleaching herbicides. In:

Toxicolody, Biochemistry and Molecular Biology of Herbicide Activity (Roe RM, Burton JD,

Kuhr, RJ), IOS Press, Amsterdam, 1-10

Sandmann G, Kowalczyk-Schröder S, Taylor HM, Böger P (1992) Quantitative structure-activity

relationship of fluridone derivatives with phytoene desaturase. Pestic Biochem Physiol 42: 1–6

Sandmann G, Linden H, Böger P (1989) Enzyme-Kinetic Studies on the Interaction of

Norflurazon with Phytoene Desaturase. Z Naturforsch 44 c: 787–790

Sandmann G, Mitchell G (2001) In vitro inhibition studies of phytoene desaturase by bleaching

ketomorpholine derivatives. J Agric Food Chem 49: 138–141

Sandmann G, Schmidt A, Linden H, Böger P (1991) Phytoene desaturase, the essential target

for bleaching herbicides. Source Weed Sci 39: 474–479

Sandmann G, Ward CE, Lo WC, Nagy JO, Böger P (1990) Bleaching herbicide flurtamone

interferes with phytoene desaturase. Plant Physiol 94: 476–478

Santabarbara S, Casazza AP, Ali K, Economou CK, Wannathong T, Zito F, Redding KE, Rappaport F, Purton S (2013) The requirement for carotenoids in the assembly and function of

the photosynthetic complexes in Chlamydomonas reinhardtii. Plant Physiol 161: 535–546

Schaub P, Yu Q, Gemmecker S, Poussin-Courmontagne P, Mailliot J, McEwen AG, Ghisla S, Al-Babili S, Cavarelli J, Beyer P (2012) On the structure and function of the phytoene desaturase

CRTI from Pantoea ananatis, a membrane-peripheral and FAD-dependent

Page 170: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

References

156

oxidase/isomerase. PLoS One (7): e39550

Schledz M, Al-Babili S, Lintig J v., Haubruck H, Rabbani S, Kleinig H, Beyer P (1996) Phytoene

synthase from Narcissus pseudonarcissus: functional expression, galactolipid requirement,

topological distribution in chromoplasts and induction during flowering. Plant J 10: 781–792

Serganov A, Nudler E (2013) A Decade of Riboswitches. Cell 152: 17–24

Shumskaya M, Wurtzel ET (2013) The carotenoid biosynthetic pathway: Thinking in all

dimensions. Plant Sci 208: 58–63

Sikkema J, De Bont JAM, Poolman B (1994) Interactions of cyclic hydrocarbons with biological

membranes. J Biol Chem 269: 8022–8028

Soetaert K, Petzoldt T, Setzer RW (2010) Solving Differential Equations in R: Package deSolve. J.

Stat. Softw. 33: 1–25

Steinbrenner J, Sandmann G (2006) Transformation of the green alga Haematococcus

pluvialis with a phytoene desaturase for accelerated astaxanthin biosynthesis. Appl Environ

Microbiol 72: 7477–7484

Suarez J V., Banks S, Thomas PG, Day A (2014) A new F131V mutation in Chlamydomonas

phytoene desaturase locates a cluster of norflurazon resistance mutations near the FAD-

binding site in 3D protein models. PLoS One 9: e99894

Takabayashi A, Kadoya R, Kuwano M, Kurihara K, Ito H, Tanaka R, Tanaka A (2013) Protein co-

migration database (PCoM-DB) for Arabidopsis thylakoids and Synechocystis cells.

Springerplus 2: 148

Tian L (2015) Recent advances in understanding carotenoid-derived signaling molecules in

regulating plant growth and development. Front Plant Sci 6: 1–4

Tian L, Dellapenna D (2004) Progress in understanding the origin and functions of carotenoid

hydroxylases in plants. Arch Biochem Biophys 430: 22–29

Wagner T, Windhövel U, Römer S (2002) Transformation of tobacco with a mutated

cyanobacterial phytoene desaturase gene confers resistance to bleaching herbicides. Z

Naturforsch 57c: 671-679

Walter MH, Strack D (2011) Carotenoids and their cleavage products: biosynthesis and

functions. Nat Prod Rep 28: 663–692

Welsch R, Beyer P, Hugueney P, Kleinig H, von Lintig J (2000) Regulation and activation of

phytoene synthase, a key enzyme in carotenoid biosynthesis, during photomorphogenesis.

Planta 211: 846–854

Wessel D, Flügge UI (1984) A Method for the Quantitaive Recovery of Protein in Dilute Solution

in the Presence of Detergents and Lipds. Anal Biochem 138: 141–143

Wheeldon I, Minteer SD, Banta S, Barton SC, Atanassov P, Sigman M (2016) Substrate

channelling as an approach to cascade reactions. Nat Chem 8: 299–309

Yu Q, Beyer P (2012) Reaction specificities of the ε-ionone-forming lycopene cyclase from rice

(Oryza sativa) elucidated in vitro. FEBS Lett 586: 3415–3420

Yu Q, Ghisla S, Hirschberg J, Mann V, Beyer P (2011) Plant carotene cis-trans isomerase

CRTISO: A new member of the FADred-dependent flavoproteins catalyzing non-redox

reactions. J Biol Chem 286: 8666–8676

Yu Q, Schaub P, Ghisla S, Al-Babili S, Krieger-Liszkay A, Beyer P (2010) The lycopene cyclase

CrtY from Pantoea ananatis (formerly Erwinia uredovora) catalyzes an FADred-dependent

non-redox reaction. J Biol Chem 285: 12109–12120

Page 171: The Poly-Cis Pathway of Carotene Desaturation: Enzymology

Acknowledgements

157

8 Acknowledgements

I would like to thank Prof. Dr. Peter Beyer for supervising my dissertation,

offering guidance and encouraging me to work and think independently.

I’m indebted to Dr. Patrick Schaub, Dr. Ralf Welsch and Dr. Florian Wüst, for

teaching me the knowledge and skills that were crucial for the success of this

work and for their technical assistance and useful advice.

I’m glad to have conducted this dissertation alongside Dr. Mark Bruno and Dr.

Daniel Álvarez and would like to thank them for their constant support. Having

spent a lot of time together, inside and outside the laboratory, I consider you

not only colleagues but good friends.

I would like to thank Carmen Schubert for her help whenever it was needed.

Thank you to Prof. Dr. Sandro Ghisla for helpful advice and providing a better

understanding of enzyme kinetics and mechanisms over the past years.

I would like to show appreciation for the work of my cooperation partners

Anton Brausemann and Dr. Mirjam Fehling-Kaschek who provided invaluable

structural and functional insight into phytoene desaturase.

A big thanks goes to Domi, Frede and Steve for sharing my passion for sports

on all our skiing and biking trips and for giving me an understanding of a new

take on things that helped me during this thesis.

Lastly but most importantly, I would like to express my deep gratitude to my

parents, my sister and my grandparents. I know you won’t agree but all I am

and have achieved, I owe to you. Thank you for always being there!