• Keine Ergebnisse gefunden

Molecular Mechanisms of Light Stress Protection in the Diatom Phaeodactylum tricornutum

N/A
N/A
Protected

Academic year: 2022

Aktie "Molecular Mechanisms of Light Stress Protection in the Diatom Phaeodactylum tricornutum"

Copied!
180
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Molecular Mechanisms of Light Stress Protection in the Diatom Phaeodactylum

tricornutum

Dissertation

zur Erlangung des akademischen Grades des Doktors der Naturwissenschaften (Dr. rer. nat.) an der Universität Konstanz, Fachbereich Biologie

vorgelegt von Sabine Sturm

Tag der mündlichen Prüfung: 18. März 2010

Referent: Prof. Dr. Peter G. Kroth

Referentin: Prof. Dr. Iwona Adamska

Konstanzer Online-Publikations-System (KOPS) URN: http://nbn-resolving.de/urn:nbn:de:bsz:352-opus-116417

URL: http://kops.ub.uni-konstanz.de/volltexte/2010/11641/

(2)
(3)

Molekulare Mechanismen der Lichtstress-Protektion in der Kieselalge Phaeodactylum tricornutum

Kieselalgen sind einzellige, eukaryotische Mikroalgen und gehören zu den am häufigsten auftretenden phytoplanktonischen Organismen auf der Erde. Sie sind photoautotroph und benutzen Lichtenergie um Photosynthese zu betreiben. Es wird geschätzt, dass sie etwa 20 % des weltweiten Kohlendioxids fixieren. Ein Überschuss an Licht, kann jedoch für die photosynthetischen Organismen schädlich sein. Es wird vermutet, dass Kieselalgen vor allem aufgrund ihrer hohen photosynthetischen Flexibilität, d.h. durch ihre Fähigkeit sich unter- schiedlichen Lichtmengen sehr schnell anzupassen, ökologisch so erfolgreich sind.

In dieser Arbeit wurden Mechanismen der Lichtstress-Protektion in der KieselalgePhaeodac- tylum tricornutummit Hilfe von molekularen Techniken in Kombination mit physiologischen Messungen untersucht. Es wurden genetische Mutanten erzeugt, die eine Veränderung im photosynthetischen linearen Elektronentransport zeigen und diese hinsichtlich ihrer Licht- anpassung charakterisiert. Die Ergebnisse zeigen, dass photoprotektive Vorgänge wie NPQ (nicht-photochemisches Quenching) oder der zyklische Elektronentransport an Photosystem II sehr wichtig sind für die Regulation der Photosynthese, um photo-oxidative Schäden am Photosynthese-Apparat zu vermeiden.

Die Bedeutung der Diadinoxanthin De-Epoxidase (DDE) für die NPQ-Entwicklung in P. tricornutum wurde untersucht. Das Enzym katalysiert die Umwandlung von Diadino- xanthin in Diatoxanthin; für letzteres konnte eine direkte Beziehung mit der möglichen NPQ Kapazität der Zelle gezeigt werden. DasDdeGen, welches für die DDE kodiert, wurde sowohl mittels RNAi gerichtetem „Gen-Silencing“ in P. tricornutum ausgeschaltet wie in den Zelle überexprimiert. Die Stilllegung der DDE verdeutlichte nicht nur die starke Be- deutung dieses Enzyms für NPQ, sondern ergab auch erste Einblicke in die Mechanismen des „Gen-Silencing“ in Kieselalgen. So lassen die vorliegenden Daten vermuten, dass zwei verschiedene Mechanismen in Kieselalgen vorhanden sind. Es konnte auch gezeigt werden, dass die Gegenwart von Diatoxanthin in ausreichender Menge nicht zwangsläufig zu höheren NPQ-Werten führt.

Die Expressionsraten von einigen Antennenproteinen wurde unter verschiedenen Lichtbe- dingungen hinsichtlich einer möglichen Funktion für die Photoprotektion untersucht. Die Er- gebnisse deuten darauf hin, dass Mitglieder der LHC-ähnlichen Familie im Gegensatz zu den entsprechenden Proteinen in höheren Pflanzen und Grünalgen nicht an der Anpassung an veränderte Lichtbedingungen beteiligt sind, jedoch bei der Lichtaufnahme mitwirken könn- ten. Auf der anderen Seite konnte gezeigt werden, dass die Transkriptsraten von drei der vier LHCX Proteine in P. tricornutum eindeutig durch Starklicht induziert werden können, was einen Hinweis für die mögliche Beteiligung an der Starklicht Anpassung in Kieselalgen darstellt.

(4)
(5)

Molecular Mechanisms of Light Stress Protection in the Diatom Phaeodactylum tricornutum

Diatoms are unicellular, eukaryotic microalgae that belong to the most abundant phyto- plankton on earth. They are photoautotrophic and use light energy to drive photosynthesis, thereby fixing an estimated amount of about 20 % of the global carbon. However, an excess of light can be harmful for the photosynthetic organisms and the ecological success of dia- toms can be partially explained by their high photosynthetic flexibility which allows them to quickly adapt to a rapidly changing light climate.

In this work, molecular tools in combination with physiological measurements were used to investigate the mechanisms of light stress protection in the diatom P. tricornutum. We produced targeted genetic mutants that are altered in the photosynthetic linear electron transport and characterized them with respect to their photoacclimative capacities. The re- sults show that photoprotective mechanisms such as NPQ (non-photochemical quenching) or the cycling electron transport of photosystem II are important to regulate photosynthesis by preventing photo-oxidative damage to the photosynthetic apparatus.

By a reverse-genetic approach, we analyzed the importance of the diadinoxanthin de- epoxidase (DDE) for NPQ in P. tricornutum. DDE catalyzes the conversion of diadino- xanthin to diatoxanthin, the latter was shown to be directly correlated with the capacity of the cells to develop NPQ. We knock-downed the respective gene encoding this enzyme by RNAi mediated gene silencing as well as overexpressed the DDE protein in vivo. Silencing the DDE did not only reveal the importance of this enzyme for NPQ, but also gave first insights into silencing mechanisms in diatoms. The data presented here suggest the pres- ence of two different gene silencing mediating mechanisms in diatoms. We also show that the presence of diatoxanthin in sufficient amounts does not necessarily lead to higher NPQ values.

Finally, we examined the expression levels of several antenna proteins under different light conditions and discuss their possible role in photoprotection. Our data indicate that, in contrast to their counterpart in higher plants and green algae, members of the LHC-like family in diatoms are not involved in photoacclimation, but might be responsible for light harvesting. Furthermore, we show that the transcript level of three of the four LHCX pro- teins inP. tricornutumare clearly induced by high light, suggesting a possible involvement in high light acclimation of diatoms.

(6)
(7)

Contents

Zusammenfassung/Abstract iii

Molekulare Mechanismen der Lichtstress-Protektion in der KieselalgeP. tricornutum iii

Molecular Mechanisms of Light Stress Protection in the Diatom P. tricornutum . . v

List of Figures xi List of Tables xiii 1. General Introduction 1 2. Characterization of psbA mutants in P. tricornutum 7 2.1. Abstract . . . 8

2.2. Introduction . . . 9

2.3. Materials and Methods . . . 10

2.3.1. Strains and media . . . 10

2.3.2. Generation of psbAmutants fromP. tricornutum . . . 10

2.3.3. Isolation of DNA and sequencing of wildtype and mutantpsbAgenes . 10 2.3.4. Cell cultivation and preparation for physiological measurements . . . . 11

2.3.5. Protein extraction and Western blot analysis . . . 11

2.3.6. Pigment extraction and analysis . . . 11

2.3.7. Spectroscopy and PSI reaction center (P700) concentration . . . 11

2.3.8. Thermoluminescence . . . 12

2.3.9. Oxygen concentration and photosynthetic light-response curves . . . . 12

2.3.10. Chl fluorescence induction kinetics and DCMU resistance . . . 12

2.3.11. Chl fluorescence yield . . . 12

2.3.12. Oxygen yield per flash . . . 13

2.4. Results . . . 14

2.4.1. Generating theP. tricornutum psbA mutants . . . 14

2.4.2. Localization of the mutations in the D1 protein and herbicide resis- tance of the P. tricornutum psbAmutants . . . 14

2.4.3. Photosystem and light-harvesting properties, and growth of theP. tri- cornutum psbAmutants . . . 15

(8)

2.4.4. Photosynthetic capacity of theP. tricornutum psbAmutants as a func-

tion of the light intensity . . . 18

2.5. Discussion . . . 19

2.5.1. A mutation that slightly affects the photosynthetic efficiency: V219I . 19 2.5.2. Effects of mutations within the QB-binding pocket: F255I and S264A 19 2.5.3. Effects of a mutation close to the nonheme iron-binding site: L275W . 20 2.5.4. Effects of similar mutations in cyanobacteria and green algae . . . 21

3. High light photoacclimation of diatom psbA mutants 23 3.1. Abstract . . . 24

3.2. Introduction . . . 25

3.3. Materials and Methods . . . 26

3.3.1. Cell cultivation and preparation for physiological measurements . . . . 26

3.3.2. Protein extraction and western-blot analysis . . . 26

3.3.3. Pigment extraction and analysis . . . 26

3.3.4. Chl fluorescence yield . . . 27

3.3.5. O2 yield per flash . . . 27

3.3.6. Transcript level analysis . . . 27

3.4. Results . . . 29

3.4.1. Response of theP. tricornutum psbAmutants to a short-term increase in light intensity . . . 29

3.4.2. Response of theP. tricornutum psbAmutants to a prolonged increase in light intensity . . . 30

3.5. Discussion . . . 37

3.5.1. Mutations in or close to the QB pocket generate impairment in NPQ ability . . . 37

3.5.2. Mutations in or close to the QB pocket modify the ability to perform the PSII CET . . . 38

3.5.3. Mutations in or close to the QB pocket generate an increased high light sensitivity . . . 39

3.5.4. How do diatoms cope with high light on a genetic and physiological level? . . . 40

4. Silencing of the DDE in P. tricornutum 45 4.1. Abstract . . . 46

4.2. Introduction . . . 47

4.3. Materials and Methods . . . 49

4.3.1. Cell cultivation and preparation for physiological measurements . . . . 49

4.3.2. PCR and construction of plasmids . . . 49

(9)

Contents

4.3.3. Biolistic transformation . . . 50

4.3.4. Isolation of RNA and cDNA synthesis . . . 50

4.3.5. Real-time PCR assays . . . 51

4.3.6. RNaseIII assays . . . 51

4.3.7. Pigment extraction and analysis . . . 51

4.3.8. Chlorophyll fluorescence yield . . . 52

4.4. Results and Discussion . . . 53

4.4.1. Silencing the Dde gene expression . . . 53

4.4.2. Screening of the Dde transformants using chlorophyll fluorescence . . 53

4.4.3. Relative quantification ofDde transcript levels via real-time qPCR . . 56

4.4.4. Pigment content, photosynthetic properties and growth rate of the WT andDde transformants ofP. tricornutum. . . 57

4.4.5. Diadinoxanthin (DD) de-epoxidation in the WT and Dde transfor- mants grown under low light . . . 58

4.4.6. Xanthophyll de-epoxidation in the WT andDdetransformants grown under ‘high light’ . . . 62

5. Investigations on DDE overexpressing transformants 65 5.1. Abstract . . . 66

5.2. Introduction . . . 67

5.3. Materials and Methods . . . 68

5.3.1. Cell cultivation . . . 68

5.3.2. PCR and construction of plasmids . . . 68

5.3.3. Biolistic transformation . . . 68

5.3.4. Isolation of RNA, cDNA synthesis and real-time PCR assays . . . 69

5.3.5. Pigment extraction and analysis . . . 69

5.3.6. Chlorophyll fluorescence yield . . . 69

5.4. Results and Discussion . . . 71

5.4.1. Transcript analysis of theDde in the WT and the transformants . . . 71

5.4.2. Growth rate and pigment content of the WT and theDde overexpress- ing transformants . . . 71

5.4.3. Chlorophyll fluorescence measurements . . . 73

6. LHC-like genes/proteins inP. tricornutum 79 6.1. Abstract . . . 80

6.2. Introduction . . . 81

6.3. Materials and Methods . . . 83

6.3.1. Cell cultivation and light experiments . . . 83

6.3.2. Sequence search and annotation . . . 83

(10)

6.3.3. Isolation of RNA, reverse transcription and quantitative PCR (qPCR) 84

6.3.4. PCR and construction of plasmids . . . 84

6.3.5. Nuclear transformation . . . 84

6.3.6. Microscopy . . . 85

6.3.7. Phylogenetic analysis . . . 85

6.4. Results . . . 86

6.4.1. Large diversity of LHC-like sequences in diatoms . . . 86

6.4.2. RedCAP sequences form conserved group encoded by one-copy genes in red lineage . . . 86

6.4.3. Localization experiments . . . 89

6.4.4. Transcript analysis upon transfer to light . . . 89

6.5. Discussion . . . 93

6.5.1. The LHC-like proteins in diatoms . . . 93

6.5.2. Expression of LHC-like and REDCAP genes inP. tricornutum . . . . 95

7. LHCX genes inPhaeodactylum tricornutum 97 7.1. Abstract . . . 98

7.2. Introduction . . . 99

7.3. Materials and Methods . . . 101

7.3.1. Cell cultivation and light experiments . . . 101

7.3.2. Isolation of RNA, reverse transcription and quantitative PCR (qPCR) 101 7.4. Results and Discussion . . . 102

8. General Discussion 105 A. Supplementary Data 109 A.1. Supplementary Material, Chapter 2 . . . 110

A.2. Supplementary Material, Chapter 3 . . . 119

A.3. Supplementary Material, Chapter 4 . . . 127

A.4. Supplementary Material, Chapter 5 . . . 129

A.5. Supplementary Material, Chapter 6 . . . 130

A.6. Supplementary Material, Chapter 7 . . . 140

B. Author Contributions 145

C. List of Publications 147

D. Acknowledgements 149

Bibliography 151

(11)

List of Figures

2.1. D1 Western Blot . . . 16

2.2. Thermoluminescence emission and fluorescence induction kinetics of WT and D1 mutants . . . 17

2.3. Chl a fluorescence parameters of WT and D1 mutants . . . 18

2.4. Influence of the mutations on the photosynthetic apparatus of the D1 mutants 21 3.1. Cyclic Electron Transfer of PSII and NPQ in the D1 mutants . . . 30

3.2. O2 evolving PSII in WT and D1 mutants . . . 31

3.3. Growth rate and Fv/Fm . . . 32

3.4. D1 WesternBlot . . . 34

3.5. Relationship between NPQ and the amount of DT+DD in the WT and the D1 mutants . . . 35

3.6. Transcript level analysis of the D1 mutants and WT of P. tricornutum . . . . 36

3.7. Summary of the reaction of WT cells and D1 mutants ofP. tricornutumupon transfer to HL . . . 42

4.1. NPQ in false color in the WT and RNAi transformants of P. tricornutum . . 54

4.2. NPQ development in the WT and the RNAi transformants of P. tricornutum 55 4.3. Relative quantification ofDde transcripts in the WT and six RNAi transfor- mants . . . 56

4.4. DEP degree and relationship between increase of DD+DT and DT amount . 59 4.5. De-epoxidation degree and DT synthesis in the WT and two RNAi transfor- mants ofP. tricornutum cells grown under LL . . . 61

5.1. Dde transcript level in WT and Dde overexpressing transformants . . . 71

5.2. Growth curves of the WT and the Dde overexpressing transformants . . . 72

5.3. De-epoxidation degree of the WT and the Dde overexpressing transformants . 74 5.4. NPQ measurements of the WT and the Dde overexpressing transformants . . 76

6.1. Sequence Alignments . . . 87

6.2. Phylogenetic Analysis . . . 88

6.3. Localization of OHP2 and REDCAP . . . 90

6.4. Transcript Level of LHC-like genes . . . 91

(12)

6.5. Gene-transfers of LHC-like and REDCAP genes during Chromealveolate evo-

lution . . . 94

7.1. Transcript level of the Lhcx genes under various light conditions . . . 103

A.1. Construction of the pGEM-T D1 transformation vectors . . . 112

A.2. Multiple sequence alignments and phylogenetic reconstruction . . . 113

A.3. Pair-wise comparison chart of D1 (psbA) amino acid similarities and distances 118 A.4. Growth curves of the WT and the D1 mutants . . . 118

A.5. NPQ in the wildtype (WT) and the D1 mutants acclimated to 250 µmol· photons·m−2·s−1 for 72 h . . . 120

A.6. Influence of the mutations on the photosynthetic apparatus and the amplitude of photoprotective mechanisms under LL and HL light acclimation in the D1 mutants . . . 121

A.7. Details on the kinetics of the PSII electron cycle (PSII CET) and NPQ in the WT and the fourpsbAmutants . . . 122

A.8. Growth and maximum PSII quantum yield for photochemistry of the WT and the four D1 mutants during acclimation to HL . . . 125

A.9. NPQ measurement at 250 µmol·photons·m−2·s−1 for the wildtype (WT) and the D1 mutants . . . 125

A.10.Relationship between NPQ and the amount of DT . . . 126

A.11.Transcript level of photosynthetic genes in WT cells grown in HL relative to LL grown cells . . . 126

A.12.LHC-like and REDCAP gene annotation table . . . 131

A.13.In silico predictions of TMH using DAS algorithm . . . 134

A.14.Hierarchical clustering of PtRedCAP expression data . . . 139

(13)

List of Tables

3.1. Photosynthetic activity, NPQ and xanthophyll cycle operation . . . 31

3.2. Pigments in D1 mutants and wildtype cells ofP. tricornutum . . . 33

4.1. Pigments and photosynthetic properties of WT and the RNAi transformants 58 4.2. DD/DT content and de-epoxidation degree of the WT and RNAi transformants 59 4.3. Fv/Fm, growth rate, DD and DT content and DEP of the WT and the RNAi transformants . . . 63

5.1. Pigment analysis of the WT and the Dde overexpressing transformands . . . 74

6.1. Summary of the transcript analysis of the LHC-like genes . . . 95

A.1. Pigment composition and photosynthetic properties of the wildtype (WT) and the D1 mutants ofP. tricornutum . . . 110

A.2. Primers used for qPCR analysis of D1 mutants and WT cells . . . 119

A.3. DD and DT contents in the WT and the D1 mutants acclimated to 250µmol· photons·m−2·s−1 for 72 h . . . 120

A.4. Primers used for the construction of silencing transformation vectors . . . 127

A.5. Relative quantification of Dde transcripts via real-time PCR . . . 128

A.6. Transcript analysis of WT and Ddeoverexpressing transformants as given by REST . . . 129

A.7. Primers used for qPCR analysis . . . 130

A.8. List of sequences used for phylogenetic analysis . . . 132

A.9. LHC-like and FCPB gene expression table . . . 134

A.10.LHCX and FCPB gene expression table . . . 140

(14)
(15)

1. Molecular Mechanisms of Light Stress Protection in the Diatom P. tricornutum, General Introduction

Sabine Sturm*

Fachbereich Biologie, Universität Konstanz, 78457 Konstanz, Germany

*E-mail: sabine.sturm@uni-konstanz.de

(16)

Diatoms and photosynthesis

Oxygenic photosynthesis is one of the most important biological processes on earth. Light energy is used by photosynthetic organisms to convert carbon dioxide (CO2) into organic compounds [93]. Although terrestrial plants make up the vast majority of photosynthetic biomass on the planet, marine phytoplankton carry out almost half of the global net pho- tosynthesis [62]. About 40 % of the organic carbon in the sea is produced by diatoms, unicellular, eukaryotic organisms that are found in most aquatic and some terrestrial habi- tats [204, 221]. Diatom photosynthesis in the sea generates about as much organic carbon as all the terrestrial rainforests combined [62]. But unlike much of the carbon generated by trees, the organic carbon produced by diatoms is consumed rapidly and serves as a basis for marine food webs. In coastal waters, diatoms support our most productive fisheries. In the open ocean, a relatively large proportion of diatom organic matter sinks rapidly from the surface, becoming food for deep-water organisms [207]. A small fraction of this sinking organic matter escapes consumption and settles on the sea floor, where it is sequestered over geological timescales in sediments and rocks and contributes to petroleum reserves.

Given the crucial role of diatoms in the global carbon cycle, plans have been made to reduce atmospheric levels of the greenhouse gas carbon dioxide by fertilizing large regions of the ocean with iron to generate huge blooms of diatoms [44]. Iron-enrichment experiments so far confirm that iron fertilization does produce the expected diatom blooms [13], but show also that most of the organic carbon generated by the bloom is consumed and recycled in surface waters. There is only a relatively small increase in the amount of cells that sink to deep waters [26]. Even large-scale fertilization projects can be expected to draw down just a small fraction of the increasing amounts of CO2 entering the atmosphere, moreover they may also have the drawback to shift community compositions and generate other greenhouse gases [44]. Nevertheless, it is important to understand how climate change may influence life in the ocean and the phytoplankton communities living in it. Diatoms do not only play an important role in the global carbon cycle, but also in the biogeochemical cycles of silica (which is part of their cell wall) and of nitrogen [207, 210, 225]. They are the most diverse group of phytoplankton with an estimated 200,000 species worldwide [154] and tend to dominate in well-mixed coastal and up-welling regions where sufficient light, inorganic nitrogen, phosphorus, silica and trace elements are available to sustain their growth [165].

Diatoms have a complex evolutionary history that is distinct from plants. Oxygenic pho- tosynthesis has its origin in cyanobacteria, while different endosymbiotic events gave rise to the photosynthetic eukaryotes like plants and algae [13] (see also figure 6.5). In an initial en- dosymbiosis a eukaryotic heterotrophic cell engulfed (or was invaded by) a cyanobacterium to form the photosynthetic plastid, red and green algae [238]. In a second endosymbiotic event a different eukaryotic heterotroph engulfed an ancestor of modern red algae. Over time, the red algal endosymbiont was transformed into plastids. The respective phylogenetic

(17)

group, the Stramenopiles, comprise for instance diatoms, brown macroalgae and plant para- sites. Due to their different evolutionary history, plastids from diatoms differ from those of higher plants and green algae. Four membranes surround the plastids of diatoms and other heterokont algae [70], while chloroplasts from land plants, green algae, red algae and glauco- phytes are enclosed by two membranes. The organization of the photosynthetic apparatus in diatoms also differs from that of higher plants. In higher plants and green algae chloro- phyll a (Chla) and b(Chlb) are the accessory pigments for light harvesting. Furthermore, diatoms are brownish in color due to the predominance of the accessory carotenoid fucoxan- thin [123], which is located together with Chl a and Chl c in their plastids as major light harvesting pigments. Fucoxanthin and chlorophylls are bound within the light harvesting complexes by fucoxanthin-chlorophyll a/c-binding proteins (FCPs) which are homologous to the chlorophylla/b-binding proteins of green algae and higher plants [77].

Photoprotection in diatoms

Although light is necessary to drive oxygenic photosynthesis, too much light can be harmful for photosynthetic organisms as it can lead to damage of all kinds of cell components such as proteins, lipids and DNA, for example by formation of highly reactive oxygen species (ROS). The decrease in the efficiency and/or the maximum rate of photosynthesis by photo- oxidative damage is called ‘photoinhibition’ [126]. Photosynthetic organisms have developed multiple photoprotective mechanisms to cope with the potentially damaging effects of light.

One of the most important mechanisms is the thermal dissipation of excess light energy, also called non-photochemical chlorophyll fluorescence quenching (NPQ). When light is absorbed by the antenna pigments, it is subsequently transferred to a Chl molecule. The resulting singlet-excitation state of Chl a (1Chl) can undergo one of four pathways. It can be re- emitted as light (chlorophyll fluorescence), it can be used to drive the electron transport chain for photosynthesis (photochemistry), it can decay through triplet chlorophyll (3Chl) and it can dissipate as heat (NPQ) [169]. The role of NPQ is to reallocate light energy that is in excess and cannot be used for photochemistry by reducing the lifetime of 1Chl and thereby preventing the production of reactive oxygen species through 3Chl. NPQ can be divided into three components [169]:

1. qE, the energy dependent quenching which is regulated by the built-up of a transthy- lakoid ∆pH and the operation of the so-called xanthophyll cycle,

2. qT, the state-transition quenching which allows reallocation of part of the energy absorbed form the photosystem II (PSII) to the photosystem I (PSI) and

3. qI , the photoinhibitory quenching.

As there is no experimental evidence for the existence of qT in diatoms [186] and qI has not been investigated so far [133], the focus will be on the ∆pH dependent quenching qE. qE is

(18)

induced by a low pH in the thylakoid lumen which is generated by the photosynthetic elec- tron transport in light. The formation of a transthylakoidal proton gradient and subsequent acidification of the lumen has two important roles for the induction of qE: the activation of the so-called xanthophyll cycle, where an epoxidized xanthophyll is de-epoxidased [194, 58], and the protonation of lumen-exposed domains of one or more PSII proteins [103, 147]. For higher plants it has been postulated that the protonation of PSII proteins and the binding of zeaxanthin, the product of the xanthophyll cycle, leads to a conformational change of the light harvesting complex (LHC) which switches the LHC into a quenched state and leads to an efficient energy dissipation [175].

In higher plants and green algae, the xanthophyll cycle (also named violaxanthin cycle) involves the conversion of violaxanthin to zeaxanthin via antheraxanthin by the enzyme violaxanthin de-epoxidase (VDE) [58]. It has been demonstrated that the amount of zea- xanthin is strongly correlated with the qE capacity [174, 177, 214]. Analyzing anArabidopsis thaliana mutant that is deficient in qE [146] revealed that the PSBS protein, encoded by thePsbS gene, is essential for the formation of qE in addition to low pH and de-epoxidased xanthophyll such as zeaxanthin [169]. In diatoms, NPQ differs from higher plants. The xan- thophyll cycle in diatoms involves different pigments, diadinoxanthin (DD) is de-epoxidased to diatoxanthin (DT) under excess light by the diadinoxanthin de-epoxidase (DDE) [92]. In contrast to the VDE of higher plants, the DDE is already activated at higher lumenal pH and is even active at pH values of about 7 [109]. Consequently, DD de-epoxidation can already be triggered by a weak proton gradient which for instance could be induced by chlororespi- ration [108]. In diatoms, both ∆pH and DT are mandatory for the development of NPQ [136, 134]. It has been widely demonstrated that the increase of NPQ is linearly correlated with the accumulation of DT among different diatom species during high light illumination [183, 30, 137, 205] although the quenching efficiency for the same amount of DT can vary among different species. Once NPQ is formed, its extent and kinetics are solely correlated to the amount of DT and independent of the transthylakoid proton gradient. This is in con- trast to land plants, where no thermal dissipation occurs without low lumen pH [174, 74].

Interestingly, the PSBS gene is completely missing in the diatom genomes ofThalassiosira pseudonanaandPhaeodactylum tricornutum[14, 25]. Still, biochemical studies indicate that there are LHC complexes in diatoms that bind DD and DT [135, 86, 20]. Recent investi- gations suggest that homologues of the LI818-like proteins might play an equivalent role in high light acclimation as PSBS in diatoms [241].

The availability of a variety of molecular tools and the continuously increasing amounts of sequence data provide new opportunities to study diatom physiology not only by physio- logical and biochemical approaches, but also using molecular techniques. The genome data of two diatom species (P. tricornutum and T. pseudonana) are available [14, 25], together with experimental molecular tools such as genetic transformation [239, 130, 12, 197, 198],

(19)

this information can be used for investigating molecular aspects of light stress protection in diatoms. In this work, molecular tools and physiological characterizations were combined to study mechanisms of photoprotection in the diatom P. tricornutum. Mutants that are altered in the linear transport rate were produced, characterized (chapter 2) and studied concerning their capacities to adapt to high light (chapter 3). We examined the importance of the diadinoxanthin de-epoxidase by a reverse-genetic approach (chapters 4 and 5) and investigated antenna proteins and their possible role in light stress protection (chapters 6 and 7).

(20)
(21)

2. First induced plastid genome mutations in an alga with secondary plastids: psbA mutations in the diatom Phaeodactylum tricornutum

(Bacillariophyceae) reveal consequences on the regulation of photosynthesis

Arne C. Materna1, *, Sabine Sturm*, Peter G. Kroth and Johann Lavaud2, Fachbereich Biologie, Universität Konstanz, 78457 Konstanz, Germany

*These authors contributed equally to this work.

Author for correspondence. E-mail: johann.lavaud@univ-lr.fr

1present address: Alm Laboratory, Civil and Environmental Engineering, Massachusetts Insti- tute of Technology (MIT), 77 Massachusetts Ave., 48-208, Cambridge, MA 02139, USA.

2present address: UMR CNRS 6250 ‘LIENSs’, Institute for Coastal and Environmental Research, University of La Rochelle, 2 rue Olympe de Gouges, 17042 La Rochelle Cedex, France.

Citation: Materna AC, Sturm S, Kroth, Lavaud J (2009) First induced plastid genome mutations in an alga with secondary plastids: psbA mutations in the diatom Phaeodactylum tricornutum (Bacillariophyceae) reveal consequences on the regulation of photosynthesis. Journal of Phycol- ogy 45: 838-846. doi:10.1111/j.1529-8817.2009.00711.x

Submitted June 27, 2008; Accepted March 16, 2009.

(22)

2.1. Abstract

Diatoms play a crucial role in the biochemistry and ecology of most aquatic ecosystems, especially because of their high photosynthetic productivity. They often have to cope with a fluctuating light climate and a punctuated exposure to excess light, which can be harmful for photosynthesis. To gain insight into the regulation of photosynthesis in diatoms, we generated and studied mutants of the diatom Phaeodactylum tricornutum Bohlin carrying functionally altered versions of the plastidicpsbAgene encoding the D1 protein of the PSII reaction center (PSII RC). All analyzed mutants feature an amino acid substitution in the vicinity of the QB-binding pocket of D1. We characterized the photosynthetic capacity of the mutants in comparison to wildtype cells, focusing on the way they regulate their photochemistry as a function of light intensity. The results show that the mutations resulted in constitutive changes of PSII electron transport rates. The extent of the impairment varies between mutants depending on the proximity of the mutation to the QB-binding pocket and/or to the non-heme iron within the PSII RC. The effects of the mutations described here forP. tricornutumare similar to effects in cyanobacteria and green microalgae, emphasizing the conservation of the D1 protein structure among photosynthetic organisms of different evolutionary origins.

Keywords

Chlorophyll fluorescence·D1 protein·Diatom·Electron transport·Herbicide·Photosystem II·QB pocket

Abbreviations

DCMU: (3-(3,4-diclorophenyl)-1,1-dimethylurea); LHC: light-harvesting complex; OEC: oxy- gen evolving complex; PAM: pulse amplitude modulation; PQ: plastoquinone; PSII RC:

photosystem II reaction center; QA and QB: quinone A and B; WT: wildtype.

(23)

2.2. Introduction

2.2. Introduction

Diatoms (Heterokontophyta, Bacillariophyceae) are a major group of microalgae ubiquitous in all marine and freshwater ecosystems. With probably >10,000 species, their biodiver- sity is among the largest of photosynthetic organisms, just after the higher plants [155].

Diatoms are assumed to contribute to about 40 % of the aquatic primary production (i.e.,

∼20 % of the annual global production) and to play a central role in the biochemical cycles of silica (which is part of their cell wall) and nitrogen [207]. Their productivity has con- tributed largely to the structure of contemporary aquatic ecosystems [60]. In contrast to the supposed primary origin of red algae, green algae, and higher plants, diatoms originate from a secondary endosymbiotic event in which a nonphotosynthetic eukaryote probably engulfed a eukaryotic photosynthetic cell related to red algae and transformed it into a plastid [117]. This peculiar evolution has led to complex cellular functions and metabolic regulations recently highlighted by the publication of the genome of two diatom species [14, 25], Thalassiosira pseudonana and Phaeodactylum tricornutum. The complex cellular functions include aspects of photosynthesis [231], photoacclimation [133], carbon and nitro- gen metabolism [8, 129], and response to nutrient starvation [7].

As for most microalgae, the photosynthetic efficiency and productivity of diatoms strongly depend on the underwater light climate [151]. Planktonic as well as benthic diatoms tend to dominate ecosystems characterized by highly turbulent water bodies (coasts and estuaries) where they have to cope with an underwater light climate with high-frequency irradiance fluctuations coupled with large amplitudes. Depending on the rate of water mixing, diatoms can be exposed to punctual or chronic excess light, possibly generating stressful conditions that impair photosynthesis (i.e., photoinactivation/-inhibition) [150, 133]. In higher plants and cyanobacteria, the processes of PSII RC photoinactivation/-inhibition are strongly in- fluenced by the redox state of the acceptor side of PSII with quinones (QA and QB) as primary electron acceptor [223, 68].

Here we report on the generation and characterization of fourpsbAmutants of P. tricor- nutum. All mutants feature distinct amino acid exchanges in the D1 protein of PSII close to or within the QB-binding pocket. The point mutations resulted in a constitutive impairment of the PSII electron transfer in all mutants to different extents. To our knowledge, this is the first report of plastid genome mutants in an alga with secondary plastids.

(24)

2.3. Materials and Methods

2.3.1. Strains and media used for producing the psbA mutants of P. tricornutum

P. tricornutum (University of Texas Culture Collection, strain 646) was grown at 22C under continuous illumination at 50 µmol·photons·m−2·s−1 in Provasoli’s enriched f/2 seawater medium [87] using Tropic Marinr artificial seawater at a final concentration of 50 %, compared to natural seawater. When used, solid media contained 1.2 % Bacto Agar (Difco Lab., Becton Dickinson and Co., Sparks, MD, USA).

2.3.2. Generation of psbA mutants from P. tricornutum

Construction of plasmid transformation vectors: Four transformation vectors were con- structed harboring a 795 bp psbA fragment containing the QB-binding pocket. The psbA inserts of each vector carried individual point mutations leading to different substitutions of the amino acid serine encoded by thepsbA(D1) codon 264 (for details and a vector map, see Fig. A.1 on page 112 in the Supplementary Material). In addition to the nonsynonymous point mutations in codon 264, a second, silent point mutation was introduced into codon 268 (TCTtoTCA, nt position 804) without changing the encoded amino acid. The purpose of the second point mutation was to delete aBssSI restriction site, thus allowing easy RFLP screening of putative transformants.

Biolistic transformation of P. tricornutum: Transformation of P. tricornutum was per- formed using a Bio Rad Biolistic PDS-1000/He Particle Delivery System (Bio-Rad, Hercules, CA, USA) as described previously [130]. Gold particles with a diameter of 0.1 µm served as microcarriers for the DNA constructs. Bombarded cells were allowed to recover for 24h before being suspended in 1mL of sterile f/2 50 % medium. Transformants (250µL) were selected at 21C under constant illumination (35 µmol·photons·m−2·s−1) on agar plates containing 5·10−6 M DCMU herbicide (3-(3,4-diclorophenyl)-1,1-dimethylurea) and repeat- edly streaked on fresh solid selective medium to obtain full segregation of the mutation.

2.3.3. Isolation of DNA and sequencing of wildtype and mutant psbA genes Total nucleic acids from the wildtype (WT) and mutant cells were isolated via a cetyltri- methylammoniumbromide (CTAB)-based method [47]. Prior to the mutagenesis of P. tri- cornutum, the WT psbA gene and the surrounding genes were sequenced (NCBI accession no. AY864816) via primer walking (GATC, Konstanz, Germany). For the molecular char- acterization of mutants, a 795 bp fragment of thepsbA gene was amplified as described in Figure A.1, and both strands were fully sequenced (GATC, Konstanz, Germany).

(25)

2.3. Materials and Methods

2.3.4. Cell cultivation and preparation for physiological measurements

P. tricornutum WT and mutant cells were grown in 200mLsterile f/2 50 % medium [87] at 21C in airlift columns continuously flushed with sterile air. The cultures were illuminated at a light intensity of 50µmol·photons·m−2·s−1 with white fluorescent tubes (L58W/25, Universal white, OSRAM GmbH, Munich, Germany) with a 16:8 h light:dark (L:D) cycle.

Cells were harvested during the exponential phase of growth, centrifuged (Allegra 25R; Beck- man Coulter GmbH, Krefeld, Germany) at 3,000 g for 10 min, and resuspended in their culture medium to a final Chla concentration of 10µgChla·mL−1. The algae were contin- uously stirred at 21C under low continuous light. For oxygen (O2), Chl fluorescence, and thermoluminescence measurements, cells were dark-adapted 20min prior to measurement.

2.3.5. Protein extraction and Western blot analysis

Cells were harvested during the exponential phase of growth as described above and subse- quently grinded in liquid nitrogen. The homogenized cells were resuspended in preheated (60C) extraction buffer (125mM Tris/HCl [pH 6.8], 4 % [w/v] SDS, 200µM PMSF, and 100 mM DTT) and, after heat treatment, protein extracted with acetone. After wash steps, the dried protein pellet was finally resuspended in extraction buffer. Total protein was separated by SDS-PAGE. Proteins were transferred electrophoretically onto a PVDF membrane (HybondTM-P, Amersham Biosciences UK Limited, Buckinghamshire, UK) and incubated with an antiserum against D1 (Anti-PsbA global antibody, AS05 084, Agrisera, Sweden). Detection was performed using the chemoluminescence detection system from Roche Diagnostics (BM Chemiluminescence Blotting Substrate POD; Roche Diagnostics GmbH, Mannheim, Germany).

2.3.6. Pigment extraction and analysis

Chl a amount was determined by spectrophotometry using the 90 % acetone extraction method. For pigment extraction, cells were deposited on a filter and frozen in liquid nitrogen.

Pigments were extracted with a methanol:acetone (70:30, v/v) solution. Pigment analysis was performed via HPLC as previously described [135]. Cell counts were performed using a Thoma hematocytometer (LaborOptik, Friedrichsdorf, Germany).

2.3.7. Spectroscopy and PSI reaction center (P700) concentration

The absorption spectra were obtained at room temperature with a DW-2 Aminco (American Instrument Co., Jessup, MD, USA) spectrophotometer, half-bandwith 3nm, speed 2nm·s−1, OD = 0 at 750 nm, 50 % f/2 medium as a reference. P700 quantity relative to Chl a was determined as described earlier [138] with the DW-2 Aminco spectrophotometer in dual beam mode (reference at 730nm).

(26)

2.3.8. Thermoluminescence

Thermoluminescence patterns were measured with a self-made thermoluminometer following the procedure previously described [71]. Flashes were single turn-over with duration of 25µs.

Samples were adjusted to 20µg Chla·mL−1 for measurement.

2.3.9. Oxygen (O2) concentration and photosynthetic light-response (P/E) curves

O2 concentration was measured with a DW1-Clark electrode (Hansatech Ltd., Norfolk, UK) at 21C. White light of adjustable intensity (measured with a PAR-sensor, LI-185A; Li-Cor Inc., Lincoln, NE, USA) was provided by a KL-1500 quartz iodine lamp (Schott, Mainz, Germany). Cell culture samples were dark-acclimated for 20 min before measurement.

P/E curves were obtained by illuminating a 2 mL sample during 5 min at various light intensities. A new sample was used for each measurement. EK, the irradiance for saturation of photosynthetic O2 emission was estimated from P/E curves.

2.3.10. Chl fluorescence induction kinetics and DCMU resistance

Chlafluorescence induction kinetics were performed with two instruments: a PEA-fluorome- ter (Walz, Effeltrich, Germany) for short-time kinetics (up to 200 ms), which allowed a classic OIJP analysis (see table A.1 on page 110 in the Supplementary Materials for details), and a self-made ‘continuous light’ fluorometer [188] for long-time kinetics (up to 100 s).

Cells were adjusted to a concentration of 5µg Chl a·mL−1 and 20 µg Chl a·L−1 for the PEA- and the self-made fluorometers, respectively.

DCMU resistance was evaluated measuring the inhibition of the PSII activity versus increasing DCMU concentrations (see table A.1 for details). The kill curves with DCMU and atrazine were performed by growing the cells on solid medium with increasing concentrations of herbicides; growth conditions were the same as described before.

2.3.11. Chl fluorescence yield

Chl fluorescence yield was monitored with a modified PAM-101 fluorometer (Walz) as de- scribed previously [138]. For each experiment, 2 mL was used. Sodium bicarbonate was added at a concentration of 4mM to prevent any limitation of the photosynthetic rate by carbon supply. When used, DCMU was incubated with the cell suspension at the begin- ning of the dark-adaptation period. Fluorescence parameters were defined as described in table A.1. The parameter used to estimate the fraction of reduced QA[27] was 1−qP where qP is the photochemical quenching of Chl fluorescence. The rate of linear electron transport

(27)

2.3. Materials and Methods

was calculated as follows:

ET R= ΦP SII×P F D×α×0.5 (2.1)

where ΦPSII is the PSII quantum yield for photochemistry, PFD is the irradiance, and αis the PSII antenna size (equivalent to 1/I1/2 ofYSS, see below).

2.3.12. Oxygen (O2) yield per flash

The relative O2 yield produced per flash during a sequence of single-turnover saturating flashes (O2 sequence) was measured polarographically at 21C with a flash electrode as described by [140]. The flashes were separated by 500 ms allowing the reopening of PSII RCs by reoxidation of QA) between each flash. The procedures used to record and calculate the steady-state O2 yield per flash (YSS, an evaluation of the number of O2 producing PSII RCs relative to Chl a), the reciprocal of the half-saturating flash intensity of flash O2-evolution saturation curves (1/I1/2 ofYSS, an evaluation of the PSII antenna size), and the miss probability per PSII were the ones described in table A.1.

(28)

2.4. Results

2.4.1. Generating the P. tricornutum psbA mutants

In an attempt to establish stable plastid transformation in P. tricornutum, we aimed for allele replacement via homologous recombination. To minimize impact on the diatom, we decided to substitute the WTpsbAgene with slightly modified versions carrying alternative point mutations in codon 264 (Fig. A.2B on page 113, green boxes, in the Supplementary Material). These mutations lead to single amino acid substitutions that were previously reported to induce herbicide resistance and a reduction in the electron transport within the PSII RC [182, 181]. Sequencing of the target region in several putative transformants revealed a variety of nonsynonymous (and in some cases additional synonymous) point mu- tations. In all experiments, the observed point mutations occurred apparently random and independently of the respective vector sequence. None of the obtained resistant strains car- ried the same pair of point mutations that was supposed to be introduced intopsbAby the utilized transformation vector (data not shown). Negative control experiments involving exclusive selection without preceding transformation, and biolistic transformation without vector DNA failed to generate resistant colonies. We sequenced 1,000 bp regions surround- ing the QB pocket as well as coding and noncoding areas more distant to the psbA locus without finding other mutations than the ones described here. Yet, we cannot exclude the possibility that additional mutations have occurred at unknown loci. However, due to the selection on DCMU, which specifically interacts with the QB pocket of the D1 protein, ad- ditional mutations at other loci are likely to be detrimental and therefore selected against.

Although intriguing, this study is not focusing on the underlying molecular mechanism leading to the elevated mutation rates in thepsbAgene; it will be the focus of a subsequent work. Instead, we characterized and compared in four selected mutants the physiological ef- fects of different amino acid substitutions in the D1 protein of the PSII RC on the regulation of photosynthesis.

2.4.2. Localization of the mutations in the D1 protein and herbicide resistance of the P. tricornutum psbA mutants

The highly conserved D1 protein is part of the PSII RC in cyanobacteria and all phyla of plastid containing photosynthetic eukaryotes (Fig. A.2A, page A.2). The QB-binding pocket is located between the DE helix and the transmembrane E helix of the D1 protein [118]. The functional relevance of the QB-binding pocket (Fig. A.2B) is highlighted by an amino acid sequence similarity of 97 %–98 % between pennate and centric diatoms, and a similarity of ∼ 90 %–93 % between diatoms and members of the red lineage, the green lineage, and even cyanobacteria (Fig. A.3, page 118). Sequencing thepsbAgenes of the four mutant strains revealed point mutations within or near the QB-binding pocket (Fig. A.2B,

(29)

2.4. Results

red squares). The mutant V219I featured an amino acid exchange (Val to Ile at position 219) in transmembrane helix D. In mutant F255I, a Phe was changed to Ile in helix DE close to the QB pocket. S264A carries a Ser to Ala substitution within the QB pocket, and in L275W, Leu was changed to Trp in the helix E.

In comparison to the WT, the competitive binding of the herbicide DCMU to the QB pocket was altered to a different degree in all mutants (table A.1), among which S264A showed the highest resistance (3,000-fold). The level of resistance was confirmed by growth curves in the presence of increasing concentrations of DCMU (not shown). S264A was also highly resistant against the herbicide atrazin (500-fold).

2.4.3. Photosystem and light-harvesting properties, and growth of the P. tricornutum psbA mutants

At low light intensity (50µmol·photons·m−2·s−1), the pigment contents of all the mutants and the WT cells were very similar, although the mutants tended to accumulate slightly more Chl a per cell (see table A.1). The concentrations of active PSII RCs per Chla,YSS and RC/CS0 were higher in all the mutants but L275W (table A.1). The low concentration for L275W was confirmed via Western blot analysis (figure 2.1). The molar PSI:PSII ratio was similar in WT and mutants with the exception of L275W, for which the ratio was higher (×1.3) ratio. The PSII LHC (lightharvesting complex) antenna size (1/I1/2 of YSS) as well as EK, the light intensity for saturation of photosynthesis, were lower in all the mutants, with the exception of V219I (table A.1).

We compared the physiological effects of the four mutations by measuring thermolumi- nescence, flash oxygen (O2) yield emission (O2 sequence), and Chla fluorescence induction kinetics. WT cells showed the expected thermoluminescence pattern with a strong B band (figure 2.2A) [51]. While V219I showed the same pattern, in F255I and S264A the tem- perature of the maximal signal was shifted from 22C to about 7C and had significantly lowered amplitude (Fig. 2.2A). The O2 sequences were highly damped in dark-adapted cells of F255I, S264A, and L275W (Fig. 2.2B) due to an increase in the miss probability (table A.1). In addition, in S264A and to a lesser extent in F255I (not shown), the O2 production was increased at flash no. 2 (due to an increase of 10 %–20 % of the S1 dark state in S264A compared to the WT), while in L275W, the maximum was at the flash no.

4 instead of no. 3. Chla fluorescence induction kinetics are shown in figure 2.2, Cand D.

All the mutants showed higher J (QAQB/QAQB state) and lower I (QAQB2− state) phases (Fig. 2.2Cand table A.1), reflecting an impairment of the QA-QB electron transfer.

The phenotype of V219I was the closest to WT phenotype, while F255I and L275W showed a significantly higher J phase (+23 % and 57 %, respectively). S264A showed a drastically increased (by 71 %) and delayed J phase and, in contrast to the other mutants, an increased I phase (see inset Fig. 2.2Cand table A.1). When recorded over a longer timescale (100 s)

(30)

Figure 2.1.:Western Blot of the D1 protein of the PSII reaction center ofPhaeodactylum tricornutumwildtype (WT) and thepsbAmutant L275W cells. Cells were grown at 50µmol·photons·m−2·s−1. Bands representing D1 degradation products of23kDaand the cross-link products of83kDaalso resulting from D1 degradation [107] were found in larger amount in L275W but not in the WT.

and at continuous illumination, the pattern of the fluorescence induction kinetics of S264A and L275W was different (only L275W is shown, Fig. 2.2D). In S264A and L275W, the am- plitude of the I-45ms peak increased, and the whole pattern of the kinetics was disturbed.

The F0 Chl a fluorescence level was increased in all mutants (table A.1). Adding DCMU (resulting in inhibition of electron transport between QAand QB) to WT cells resulted in an increasedF0(195±6.5) comparable to S264A and L275W. When grown at low light inten- sity (50µmol·photons·m−2·s−1) all mutants showed a maximum photosynthetic efficiency of PSII (Fv/Fm, table A.1), which was similar to the WT cells, except for L275W (−19 %).

When measured at an equivalent irradiance, the effective PSII quantum yield (ΦPSII, ta- ble A.1) was the same for WT cells and V219I, but lower in the other mutants. These values were in accordance with the steady-state electron transport rate per PSII (ET0/CS0, table A.1). Addition of DCMU to the WT resulted in a decreased ΦPSII (0.38), similar to that of S264A and L275W. Only L275W showed a reduction in growth rate −26 % (µ, table A.1) and final maximal biomass (Fig. A.4, page 118). Although F255I and S264A reached the same final biomass with the same growth rate as the WT, they showed a 24 h delay (see days 3 and 2, respectively, Fig. A.4).

(31)

2.4. Results

Figure 2.2.:(A) Thermoluminescence emission of dark-adapted cells of Phaeodactylum tricornutum wildtype (WT) and thepsbAmutants V219I/F255I/S264A. The characteristic emission bands at 7C (Q) and 22C (B) are shown; they reflect the recombination states of the PSII reaction center S2QA

and S2/3QB

and the redox potential of QA and QB, respectively [71, 51]. Curves represent the average of three measurements. (B) O2production in a series of single-turnover flashes (O2sequences) by dark-adapted cells ofP. tricornutumWT and of the twopsbAmutants S264A and L275W, as measured via a flash electrode. The pattern of the O2

sequence for V219I resembled the one of the WT, and the pattern of F255I resembled the one of S264A with less pronounced features. See table A.1 (in the Supplementary Material) for a detailed description. (C-D) Chlafluorescence-induction kinetics reflect quantum yield changes of Chlafluorescence as a function of the illumination duration, which relates to both excitation trapping in PSII and the ensuing photosynthetic electron transport. (C) Short-time kinetics recorded via PEA fluorometer from dark-adapted cells of P. tricornutum WT and the fourpsbAmutants (V219I/F255I/S264A/L275W). The letters O, J, I, P, H, and G refer to the phases of the kinetics [142]. (D) Long-time kinetics from dark-adapted cells of WT and L275W (same pattern for S264A) as recorded with a self-made ‘continuous light’ fluorometer. The arrow indicates the first peak (I phase at 45ms). The amplitude of I reflects the redox state of QA [134]. In diatoms, the classic P peak is divided into two peaks, H and G [134, 142].

(32)

Figure 2.3.:Chlafluorescence parameters as recorded with a PAM-fluorometer for the wildtype (WT) and the fourpsbAmutants (V219I/F255I/S264A/L275W) ofPhaeodactylum tricornutumcells as a function of a light intensity gradient from darkness (0µmol·photons·m−2·s−1) to the equivalent of full sunlight in nature (2,000µmol·photons·m−2·s−1, [150]. The illumination duration was 5min; a new sample was used for each irradiance treatment. (A) 1−qP estimates the fraction of reduced QA [27]. Inset: Ratio mutants versus WT of the amplitude of the I-45 ms peak (see Figure 2.2B) up to 100µmol·photons·m−2·s−1. (B) ETR is the rate of linear electron transport. See the Materials and Methods section (chapter 2.3) and table A.1 (in the Supplementary Materials) for details about the calculations of these parameters. Values are average±SD of three to four measurements.

2.4.4. Photosynthetic capacity of the P. tricornutum psbA mutants as a function of the light intensity

The light intensity dependent impairment of the QA-QB electron transfer was evaluated by measuring 1−qP, a fluorescence parameter that estimates the fraction of reduced QA [27].

While 1−qP was similar in WT and in V219I, it was the highest in S264A and L275W (Fig. 2.3A). A difference in the extent of QA reduction was also found at rather low light intensities (inset Fig. 2.3A) as indicated by the ratios of the extent of the I-45mspeak from the long-time fluorescence induction kinetics of mutant versus WT (see Fig. 2.2D). In S264A and L275W, 1−qP reached saturation earlier (between 250 and 400µmol·photons·m−2·s−1) than in WT cells. In F255I, the extent of QA reduction was higher than in WT up to a light intensity of 400 µmol·photons·m−2·s−1. The direct consequences of the impaired QA-QB electron transfer were changed amplitudes of the electron transport rate per PSII (ETR) as well as altered patterns of ETR as a function of light intensity (Fig. 2.3B). The maximum ETR was decreased in all the mutants but to a different extent, thus confirming the values for ET0/CS0 (table A.1). In contrast to WT and the other mutants, ETR was already maximal in S264A and L275W at a light intensity of 250µmol·photons·m−2·s−1; at this light intensity, the extent of QA reduction was close to its maximum (Fig. 2.3A).

(33)

2.5. Discussion

2.5. Discussion

Three out of the four psbAmutants showed a phenotype clearly distinct from the WT (see Fig. 2.4 on page 21). Obviously, the observed amino acid substitutions hold implications for the phenotype of the mutants. The phenotypic effects described in this study allow various insights into the functionality of mutated residues or domains within the D1 protein.

2.5.1. A mutation that slightly affects the photosynthetic efficiency: V219I In response to the slightly increased reduction state of QA(+10 %) and the decreased ETR per PSII (about 5 %), in V219I the number of PSII RCs was increased (14 % to 22 %, depending on the method) to maintain a photosynthetic activity similar to the WT as reflected also by its growth pattern. Hence, the exchange of Val to Ile at the position 219 in the helix D appears to be too distant from the QB-binding pocket to significantly disturb the electron transport within the PSII RC inP. tricornutum.

2.5.2. Effects of mutations within the QB-binding pocket: F255I and S264A The residues Phe255 and Ser264 bind the head group of QB [118]. The electron transport between QAand QBin F255I was significantly impaired (Fig. 2.4), slowing down the reoxida- tion of QAas illustrated by the increased QAQB/QAQBstate. It was especially visible with the pattern of thermoluminescence that resembles the one reported inP. tricornutum for WT cells treated with DCMU (abolishment of the B band and increase of the Q band) [51]. Backward electron transfer from QB to QA, as illustrated by an enhancedF0 in pho- tochemically inactive PSII RCs [234], might partially explain the increased concentration of QA. As a consequence, the miss probability of the S statecycle was increased, the S1 state was stabilized [191], and the lifetimes of the redox states S2 and S3increased [73], indicating a disturbed OEC operation. To compensate the decreased photochemistry of PSII, in F255I the number of PSII RCs increased (Fig. 2.4), reflected by a slight increase of Chlaper cell as also reported for higher plants [211]. Nevertheless, the overall amount of pigments per Chl a did not change; thus, the antenna size per PSII decreased, leading to a similar increase in EK. Decreasing the PSII antenna size is known to be a straightforward way to relief from high excitation pressure on PSII due to a slowed down electron flow within the PSII RC because of mutations, herbicides, environmental stress, or other factors. Similarly, it was suggested that inP. tricornutum an increased number of photosynthetic units together with decreased size of these units might allow maximization of photochemistry at different light regimes [226], which might be the case in the mutants F255I and S264A. In spite of all these changes, the potential for photochemistry, qP, was decreased at intermediary irradiances (up to 400µmol·photons·m−2·s−1). At high light intensities, the ETR was reduced (Fig. 2.4), reflecting the decrease of the PSII antenna size and of ΦPSII.

(34)

S264A showed a more drastic reaction compared to F255I regarding the QA reoxida- tion, the electron back transfer QB to QA, but also the QB/QB2− reoxidation (increased QAQB2− state) (Fig. 2.4). Consequently, the operation of the OEC S-state cycle was strongly disturbed similar to the pattern of the fluorescence induction kinetics, illustrating the consequence of the modified QA-QB redox state on the whole electron transport chain and especially on the redox state of the plastoquinone (PQ) pool [142, 187]. As F255I, S264A reacted by increasing the PSII number. qP was largely diminished, which usually reflects accumulation of dysfunctional, highly reduced PSII RCs. It led to a decrease in ETR at all light intensities. Both qP and ETR were saturated at a much lower irradiance than in the WT. The exchange of Ser to Ala probably modified the spatial arrangement of the QB pocket [73, 191], as illustrated by the high DCMU resistance, and greatly impaired not only the redox state of QB but the binding of QB itself [36].

2.5.3. Effects of a mutation close to the nonheme iron-binding site: L275W Leu275 is close to one of the histidines binding the nonheme iron atom (His272 in helix E, grey bar in Fig. A.2, page 113), as well as at nearly equal distance between QA and QB [118]. The effect of the L275W mutation on the photosynthetic ability per PSII was similar to the point mutation S264A (Fig. 2.4) but showed a highly disturbed OEC operation along with increased QA reduction. QA reduction was already elevated (1.2- to 1.3-fold compared to WT) at light intensities that were even below the intensity used for growing the cells.

Ultimately, L275W showed a decreased growth rate under low light as well as the inability to reach the same final maximal biomass. The main difference compared to the other mutants was the reduced amount of active PSII (Fig. 2.4), demonstrated by a disturbed D1 repair cycle. Mutations close to or within the QB pocket have been reported to modify the D1 turnover either by accelerating its damage and/or by inhibiting its proteolysis and/or synthesis [36, 173]. Thus, it is very likely that in L275W there is a mixed population of active and inactive PSII, even at low light intensities [162], which is supported by the highF0level and the lowest Fv/Fm. In contrast to the other mutants, L275W responded to the point mutation by modifying the architecture of the photosynthetic apparatus as illustrated by the increase of the PSI:PSII stochiometry (Fig. 2.4). This attempt to maintain a reasonable photosynthetic activity might lead to an increased capacity for PSI cyclic electron flow as in higher plants with deficient linear electron transport [127].

In a series of papers (reviewed in [222]), Govindjee et al. showed that the exchange of the residue Leu275 significantly perturbs the QA-Fe-QB structure, the protonation of QB2−

[234], and subsequently the PQ redox state. It is thus likely that the phenotype of L275W is due to the close vicinity of the point mutation to both the QB pocket and the nonheme iron atom binding sites, which functionally affects the properties of both the QA and QB pockets [224].

(35)

2.5. Discussion

Figure 2.4.:Diagram of the influence of mutations on the photosynthetic apparatus of the psbA mutants (V219I/F255I/S264A/L275W) ofPhaeodactylum tricornutum in comparison to the wildtype (WT) situation.

Left: the electron pathways within the PSII reaction center; right: the architecture of the photosystems as a function of WT situation (PSI:PSII 1:2); arrow up: increased value (the true value is given in-between brackets);

arrow down: decrease; flat arrow: no change. Symbols: red star, mutation; size of QA

/QB

, concentration of QA

/QB

; thickness of the earrows, value of the ETRmax and of the QB

to QA back-transfer; dotted feature of the OEC arrow, proportion of the disturbance of the OEC operation. EK, light intensity for satura- tion of photosynthesis; e, electrons; LHC, light-harvesting antenna complex; OEC, oxygen evolving complex;

PSI/PSII RC, PSII/PSI reaction center; PSI:PSII, molar photosystem stoichiometry; QAand QB, quinones. See the text for a more detailed description.

2.5.4. Effects of similar mutations in cyanobacteria and green algae

The effects of the mutations described here for the diatom P. tricornutum are similar to effects of the same mutations reported in other photosynthetic organisms. For example, it has also been concluded that in the green algaChlamydomonas reinhardtii [57], the V219I amino acid substitution does not significantly disturb the electron transport within the PSII RC. Also, the effects of the F255I, S264A, and L275W mutations have been described in cyanobacteria (Synechocystis and Synechococcus) and C. reinhardtii [57, 59, 73, 124, 191].

Remarkably, the S264A-induced DCMU resistance was much higher inP. tricornutum than in all previously studied organisms [73].

Referenzen

ÄHNLICHE DOKUMENTE

Results corresponding flash intensity (Fig. By comparing the different flash intensities which are required for reaching 50 % of the steady state oxygen yield, estimations bout

In order to establish a stable plastid transformation system for the diatom Phaeodactylum tricornutum based on homologous recombination we followed three different strategies

Hence, LHCX gene expression is apparently fine-tuned by the light quality (cryptochrome photoreceptor) as well as the light intensity either in situations that generate an

In this thesis the light and time dependent expression of the Calvin cycle was investigated with a special focus on the higher plants centrally regulated enzymes of the Calvin cycle

• We used RNA-interference to silence the single gene encoding pyruvate-orthophosphate dikinase (PPDK) in Phaeodactylum tricornutum, essential for C4 metabolism,

Expression of the early cell cycle marker genes cyclin H1 (CYCH1) and the transcription factor E2F1 was extended in the silenced versus wild-type cells (Figure 3D), indicating

induce or enhance the acclimation to higher light intensities. The predicted nuclear localisation of all four aureochromes and the confllination by successful GFP fusion

De-epoxidation state (DES) and diatoxanthin synthesis in the WT and two Dde transformants of Phaeodactylum tricornutum cells grown under low light as a function of time at an