university of groningen from photosystem i to photosystem

149
University of Groningen From Photosystem I to Photosystem II Drop, Bartlomiej Andrzej IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite from it. Please check the document version below. Document Version Publisher's PDF, also known as Version of record Publication date: 2014 Link to publication in University of Groningen/UMCG research database Citation for published version (APA): Drop, B. A. (2014). From Photosystem I to Photosystem II. s.n. Copyright Other than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of the author(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons). The publication may also be distributed here under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license. More information can be found on the University of Groningen website: https://www.rug.nl/library/open-access/self-archiving-pure/taverne- amendment. Take-down policy If you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediately and investigate your claim. Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons the number of authors shown on this cover page is limited to 10 maximum. Download date: 08-05-2022

Upload: others

Post on 08-May-2022

10 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: University of Groningen From Photosystem I to Photosystem

University of Groningen

From Photosystem I to Photosystem IIDrop, Bartlomiej Andrzej

IMPORTANT NOTE: You are advised to consult the publisher's version (publisher's PDF) if you wish to cite fromit. Please check the document version below.

Document VersionPublisher's PDF, also known as Version of record

Publication date:2014

Link to publication in University of Groningen/UMCG research database

Citation for published version (APA):Drop, B. A. (2014). From Photosystem I to Photosystem II. s.n.

CopyrightOther than for strictly personal use, it is not permitted to download or to forward/distribute the text or part of it without the consent of theauthor(s) and/or copyright holder(s), unless the work is under an open content license (like Creative Commons).

The publication may also be distributed here under the terms of Article 25fa of the Dutch Copyright Act, indicated by the “Taverne” license.More information can be found on the University of Groningen website: https://www.rug.nl/library/open-access/self-archiving-pure/taverne-amendment.

Take-down policyIf you believe that this document breaches copyright please contact us providing details, and we will remove access to the work immediatelyand investigate your claim.

Downloaded from the University of Groningen/UMCG research database (Pure): http://www.rug.nl/research/portal. For technical reasons thenumber of authors shown on this cover page is limited to 10 maximum.

Download date: 08-05-2022

Page 2: University of Groningen From Photosystem I to Photosystem

1

From Photosystem I to Photosystem II Travelling across the thylakoid membrane of

Chlamydomonas reinhardtii with LHCII

Bartlomiej Drop

Page 3: University of Groningen From Photosystem I to Photosystem

2

This Phd study was carried out at the Groningen Biomolecular Sciences and Biotechnology

Institute, Faculty of Mathematics and Natura Sciences, University of Groningen.

ISBN: 978-90-367-6841-2

ISBN (electronical version): 978-90-367-6840-5

Cover design by Marija Smits

Printed by Ipskamp Drukkers, Enschede

Page 4: University of Groningen From Photosystem I to Photosystem

3

From Photosystem I to Photosystem II

Travelling across the thylakoid membrane of Chlamydomonas reinhardtii with LHCII

PhD thesis

to obtain the degree of PhD at the University of Groningen on the authority of the

Rector Magnificus Prof. E. Sterken and in accordance with

the decision by the College of Deans.

This thesis will be defended in public on

Friday 14 March 2014 at 16.15 hours

by

Bartłomiej Andrzej Drop

born on 17 February 1983 in Czeladź, Poland

Page 5: University of Groningen From Photosystem I to Photosystem

4

Supervisor Prof. dr. R. Croce Assessment committee Prof. dr. E.J. Boekema Prof. dr. S.J. Marrink Prof. dr. B. Robert

Page 6: University of Groningen From Photosystem I to Photosystem

5

Contents

Chapter 1 General introduction 7

Chapter 2 Photosystem I of Chlamydomonas reinhardtii is composed of nine

Light-harvesting complexes (Lhca) located on one side of the core

41

Chapter 3 Light-harvesting complex II (LHCII) and its supramolecular organization in

Chlamydomonas reinhardtii

69

Chapter 4 Consequences of state transitions on the structural and functional

organization of Photosystem I in the green alga Chlamydomonas

reinhardtii

95

Chapter 5 During state 1 to state 2 transition in Arabidopsis thaliana the Photosystem

II supercomplex gets phosphorylated but does not disassemble

121

Summary 139

Samenvatting 143

Acknowledgments 147

Page 7: University of Groningen From Photosystem I to Photosystem

6

Page 8: University of Groningen From Photosystem I to Photosystem

7

CHAPTER 1

General introduction

Page 9: University of Groningen From Photosystem I to Photosystem

8

Abstract

The purpose of this thesis is to provide a comprehensive picture of the structural and

functional properties of Photosystem I (PSI) and Photosystem II (PSII) supercomplexes of the

green alga Chlamydomonas reinhardtii (C.r.). This chapter presents an overview of the topics

addressed in this thesis. The essential characteristics and significance of Chlamydomonas

reinhardtii are introduced, as well as the the evolution of photosynthetic organisms and

principles of the light reactions of photosynthesis. Next the most recent data about

structural and functional properties of the photosynthetic complexes belonging both to

green alga and higher plants, with emphasis on C. reinhardtii PSI and PSII, are discussed.

Finally, the focus will be on the reorganization of photosynthetic complexes within the

thylakoid membranes in particular and the rearrangement of the core and light harvesting

antenna that occurs during photoacclimation processes – high light acclimation and state

transitions.

Page 10: University of Groningen From Photosystem I to Photosystem

9

1. Green alga Chlamydomonas reinhardtii

The research presented in this thesis was mainly carried on a unicellular green alga,

Chlamydomonas reinhardtii. The genus Chlamydomonas can be found all over the world, i.e.

in freshwater lakes, sewage ponds, snow, garden and agricultural soil, forests, deserts, peat

bogs, damp walls, even in mattress dust in the Netherlands or roof tiles in India (Harris

2008).

From all Chlamydomonas species, C. reinhardtii is the most widely used in laboratory

research. The principal laboratory strains of C. reinhardtii were isolated by GM Smith in

1945 from soil collected near Amherst, Massachusetts (Harris 2001).

1.1. The biology of Chlamydomonas reinhardtii

Historically, Chlamydomonas species were defined only based on morphological

criteria. C. reinhardtii cell is oval shaped, approximately 10 μm in length and 3 μm in width.

There are two flagella (12–14 µm) at their anterior end that are used to swim and sense

environmental conditions. The cell contains: a single cup-shaped chloroplast that occupies

nearly 40% of its volume, a nucleus with a large nucleolus, several mitochondria,

an endoplasmatic reticulum, the Golgi apparatus, and contractile vacuoles. Like higher

plants C. reinhardtii has a cell wall that, together with vacuoles, maintains the osmotic

balance of the cell and prevents lysis of the plasma membrane. Within the C.r. chloroplast

are: the pyrenoid that is used for starch storage, and the eyespot, which is composed of

several layers of pigment granules (rhodopsin-family photoreceptors) that is used to detect

light and trigger phototaxic responses. The vegetative cell is haploid and multiplies through

mitotic divisions. The vegetative cell differentiates into gametes that exist in two mating

types, mt+ and mt- (Rochaix 2001). Cells of opposite mating type fuse and form a zygote,

which then undergoes meiosis and form four haploid daughter cells that resume vegetative

growth (Harris 2001; Rochaix 2001; Harris 2008).

C. reinhardtii cells may be cultivated under three different growing conditions: (1) minimal

medium with light and CO2 as the sole carbon source (phototrophic growth), (2) acetate

containing medium with light (mixotrophic growth), or (3) without light (heterotrophic

growth) (Rochaix 2001).

Page 11: University of Groningen From Photosystem I to Photosystem

10

1.2. Chlamydomonas reinhardtii as a model organism

The completely sequenced genome, a vast collection of mutants, simple life cycle

and easy maintenance are the reasons why Chlamydomonas reinhardtii is an excellent

organism to work with, attracting more and more attention in scientific society. This green

alga is used to study a variety of cellular activities, including mechanisms of photosynthesis

(Rochaix 2002). In recent years, C. reinhardtii has also become an important candidate for

industrial applications, such as bioethanol or biohydrogen production (Melis, Zhang et al.

2000; Rupprecht 2009).

C. reinhardtii combines the advantages of a simple unicellular organism with a sophisticated

eukaryotic system of higher plants. In contrast to higher plants, the cells can grow also

heterotrophically (Rochaix 2001). In addition, C. reinhardtii is able to synthesize

photosynthetic pigments in both light-dependent and light-independent pathways, and as a

consequence has fully assembled photosynthetic complexes even in the dark. This feature

allows the identification of mutants which are lacking photosynthetic function (Davies and

Grossman 1998; Rochaix 2001). Both the nuclear and the chloroplast transformations

of C. reinhardtii are easy to carry out and, unlike in higher plants, the time from the initial

transformation to final result is relatively short (Mayfield and Franklin 2005). Another

advantage of C. reinhardtii is that cells can readily uptake compounds that serve as electron

donors or electron acceptors in the photosynthetic electron transport chain, or inhibitors

that block photosynthesis. This green alga is also suitable for in vitro studies, because the

cost of growing C. reinhardtii is low and large amounts of cells can be grown quickly

(Mayfield and Franklin 2005).

2. Photosynthesis

Photosynthesis supplies Earth’s biosphere with oxygen and energy for living

organisms. In this process, the energy of sunlight is absorbed and converted into chemical

energy. At the heart of photosynthesis is the water splitting reaction. The water oxidation

results in the formation of molecular oxygen and reducing equivalents, which are used for

the fixation of carbon dioxide to organic matter (Arnon 1971; Barber 2002).

6H2O + 6 CO2 → (CH2O)6 + 6O2

Page 12: University of Groningen From Photosystem I to Photosystem

11

Photosynthetic O2 production and carbon dioxide assimilation established the composition

of the biosphere that led to the development of advanced life forms (Xiong and Bauer

2002).

2.1. Evolution of light harvesting organisms

Photosynthesis evolved very early in the history of life, although the earliest events

in its evolution are not clear (Nisbet and Sleep 2001). In the absence of an ozone layer, the

ancient life forms could not have survived the Sun’s harmful high-frequency radiation.

Therefore, photosynthesis must have arisen in water reservoirs under considerable aquatic

protection (Schopf 1978; Olson and Blankenship 2004). At that time, photosynthetic

pigments functioned mainly for the protection of the DNA and proteins from UV damages

(Blankenship 2010). It is hypothesized that the chlorophylls assembled together with specific

proteins to form primordial reaction centers that were adopted by the DNA/protein-based

organisms, and these eventually evolved into the different photosynthetic complexes

(Mulkidjanian, Koonin et al. 2006; Amunts and Nelson 2008) .

The evolution of photosynthesis accelerated when the photochemical reaction centers (RCs)

occurred (Mulkidjanian, Koonin et al. 2006). The first traces of carbon fixation by

photosynthetic organisms were identified in sediments formed 3.8 billion years ago

(Mojzsis, Arrhenius et al. 1996). These first photosynthetic bacteria - cyanobacteria - began

to be dominant due to the energetic advantage of photosynthesis (Blankenship 1992).

In terms of energy production, the photosynthetic process of reduction NADP+ to NADPH for

organic biosynthesis by cleaving water was much more efficient than the available

oxidation-reduction reactions (Olson 1981). This, in conjunction with constant supplies of

sunlight, moisture, and nutrients, resulted in massive evolutionary bloom of cyanobacteria.

Moreover, by the creation of the potentially toxic agent O2, cyanobacteria’s competitors

were inhibited (Blankenship and Hartman 1998). Oxygen production and accumulation

in the atmosphere lead to one of the most important events in the history of the Earth –

“the Big Bang of Evolution’’. During that period most organisms vanished and the rest had to

adapt to changes in the environment (Barber 2004). Another significant event that occurred

as a consequence of oxygen accumulation was the appearance of eukaryotic cells (Amunts

and Nelson 2009). Some eukaryotic cells phagocytized primordial cyanobacteria through

concomitant endosymbiosis, which resulted in the formation of chloroplasts, special

Page 13: University of Groningen From Photosystem I to Photosystem

12

organelles in which the process of oxygenic photosynthesis occurs (Yoon, Hackett et al.

2004; Amunts and Nelson 2008). Following this primary endosymbiosis, the photosynthetic

eukaryotic lineage diverged into three lines: (1) Rhodophyta, (2) Glaucophyta, and (3)

Viridiplantae. One billion years ago the Viridiplantae branched into the Chlorophyta (green

algae, including C. reinhardtii), and the Spermatophyta (higher plants and their close

relatives) (Merchant, Prochnik et al. 2007).

2.2. Photosynthetic reaction centers

In marine and halophilic archaea, light energy is used by bacteriorhodopsin and

halorhodopsin in photoreception-like processes to produce a transmembrane potential

difference and as a consequence to pump ions (protons and chloride ions respectively)

across a membrane (Riesle, Oesterhelt et al. 1996; Beja, Spudich et al. 2001).

In photosynthetic prokaryotes and eukaryotes, light energy is used for charge separation,

which initiates electron transport and generates a proton gradient across the membrane.

This process is driven by the photosynthetic reaction centers. Structural, spectroscopic,

thermodynamic, and molecular sequence analysis segregates all known reaction centers

into two groups called type I and type II (Fig. 1) (Blankenship 2010).

Figure 1. Electron transport diagram indicating the types of RCs: type II (purple) and type I (green), and electron transport pathways found in different groups of photosynthetic organisms (Blankenship 2010).

The primary difference between the two types of RCs is the nature of the early electron

acceptor cofactors. FeS is an electron acceptor in type I RC, while pheophytin/quinone

complexes are an electron acceptor in type II RC. Photosynthetic organisms which have only

Page 14: University of Groningen From Photosystem I to Photosystem

13

one type of RC are called anoxygenic phototrophs. These are divided into two groups:

(1) sulphur bacteria and helicobacter that have the type I RC (PSI-like) and (2) purple

bacteria and green filamentous bacteria that have the type II RC (PSII-like). Cyanobacteria,

green algae and higher plants are called oxygenic phototrophs and they all have both type I

and type II RCs (PSI and PSII respectively) that work in series to oxidize water and reduce

NADP+ (Blankenship 2010).

Phylogenetic analyses of PSI and PSII core reaction center proteins from oxygenic

phototrophs reveal high structural similarities with the type I RC of green sulfur bacteria and

type II RC of purple bacteria (Deisenhofer and Michel 1991; Schubert, Klukas et al. 1998).

These similarities support the hypothesis that all reaction center complexes derive from the

same ancestral photosystem (Blankenship and Hartman 1998).

2.3 Thylakoid membranes and main photosynthetic complexes

In oxygenic photosynthetic organisms the light reactions of photosynthesis occur in highly

specialized membranes within the chloroplast (Fig. 2A). These membranes are called

thylakoids and harbor four major photosynthetic complexes – PSI, PSII, cytochrome b6f (Cyt

b6f) and ATP synthase (ATPase). In addition, the soluble electron carriers ferredoxin (Fd) and

ferredoxin-NADP+-reductase (FNR) are present on the stromal side, and plastocyanin (PC) or

cytochrome c6 (Cyt c) act as electron carriers on the lumenal side (Staehelin 2003).

Thylakoids form a physically continuous three-dimensional network enclosing an aqueous

space called lumen. The membranes are extensively folded (Fig. 2B) and as a consequence

differentiated into two distinct physical domains: (1) cylindrical stacked structures (grana)

and (2) interconnecting single membrane regions (stroma lamellae). In green algae and

higher plants, the protein complexes that catalyze electron transfer are unevenly distributed

in thylakoid membranes: PSII is located mainly in the grana membanes, while PSI and

ATPase reside mainly in the stroma lamellae. The cytohrome b6f complex is distributed

evenly between these two domains (Fig. 2C) (Dekker and Boekema 2005; Nelson and Yocum

2006). The cooperation of these complexes results in the formation of ATP and NADPH

(Amunts, Toporik et al. 2010).

Page 15: University of Groningen From Photosystem I to Photosystem

14

Figure 2. Chloroplasts and thylakoid membranes. A: Schematic diagram of a chloroplast; B: Electron micrograph of a chloroplast (Taiz and Zeiger 2006); C: Molecular topography of thylakoid membranes and lateral heterogeneity in the distribution of PSI and PSII (Allen and Forsberg 2001).

2.4. Molecular mechanism of photosynthesis

Despite billions of years of separate evolution between cyanobacteria, green algae

and higher plants, the photosynthetic apparatus of these organisms is very similar and

in accordance photosynthesis operates with the same mechanisms.

2.4.1. Linear electron flow

In oxygenic photosynthesis, PSI and PSII work in series in the so-called linear electron

flow (LEF) or Z-scheme. PSII catalyzes the light-driven oxidation of water. This process

provides electrons to PSI via plastoquinone, Cyt b6f complex and plastocyanin (Fig. 3).

The electron transport chain (ETC) generates a transmembrane electrochemical potential

gradient that enables the ATPase to produce ATP. PSI catalyzes light-driven electron

transport from plastocyanin on the lumenal side of the membrane to ferredoxin on the

stromal side. Reduced ferredoxin is then used for NADPH production. Next, in the dark

reactions, NADPH and ATP are used to convert CO2 into organic molecules.

After absorption of a photon by antenna pigments of PSII, the excitation energy is rapidly

transferred to the PSII reaction centre – P680 (Primary electron donor absorbing at 680 nm)

(van Grondelle, Dekker et al. 1994). After excitation, P680 becomes photooxidized and

donates an electron to the primary acceptor of PSII – pheophytin. This charge separation is

then further stabilized by the transfer of the electron to plastoquinones (PQ), first to the QA

and then to the QB binding site, located close the stromal side of the PSII complex. Double

reduction of QB leads to its release from the PSII binding site as a plastoquinol (PQH2) that

diffuses toward the Cyt b6f. Before a second reduction of QB, P680+ is reduced by YZ (redox

Page 16: University of Groningen From Photosystem I to Photosystem

15

active tyrosine residue) located on the D1 protein. The subsequent reduction of YZ, oxidizes

the Mn4CaO5 cluster that cycles through several different redox states (Sn, n = 0–4). S0 is the

most reduced and S4 is the most oxidized state formed during the four-electron cycle (Kok,

Forbush et al. 1970; Radmer and Kok 1975). Upon oxidation to the S4 state, the Mn4 cluster

is spontaneously reduced to S0 by oxidizing water to form O2 and H+, which are released into

the lumen (Debus 1992; Yachandra, Sauer et al. 1996).

Figure 3. A detailed model of the electron transport chain including structural information on the organization of the protein complexes involved in electron (e-) and proton (H+) transport within the thylakoid membrane of higher plant Arabidopsis thaliana (Allen, de Paula et al. 2011).

Electron transfer between PSII and PSI is linked by the Cyt b6f, which contains cytochromes

and a Rieske iron-sulfur centre. When plastoquinol reaches Cyt b6f it is oxidized. One

electron is transferred through Rieske protein and cytochrome f to a small soluble copper

protein – plastocyanin. The second electron is transferred through two b hemes of

cytochrome b6 and is used to reduce another quinone from the PQ pool at the stromal side.

In a second reduction event, two protons are taken from the stroma and subsequently

released into the lumen, generating an electrochemical gradient across the membrane. This

recycling of one of the electrons from the PQ pool is called Q-cycle (Kurisu, Zhang et al.

2003). As a result, plastoquinone is recycled to PSII, plastocyanin diffuses through the lumen

to PSI, and protons are released into the inner thylakoids space (Nugent 1996; Nelson and

Ben-Shem 2004).

Page 17: University of Groningen From Photosystem I to Photosystem

16

As in PSII, an analogous process occurs in PSI. Light is absorbed by antenna pigments and

excitation energy is transferred to the PSI reaction centre. After primary charge separation

that is initiated by the excitation of the reaction center - P700 (Primary electron donor

absorbing at 700 nm) - the electron passes along the electron transfer chain that consists of

several cofactors: A0 (Chl a), A1 (phylloquinone), and the iron-sulphur (Fe4S4) clusters – FX, FA

and FB. At the stromal side, the electron is donated by FB to ferredoxin and then transferred

to the ferredoxin NADP+ reductase. The reaction cycle is completed when P700+ is re-

reduced by plastocyanin at the lumenal side of the membrane (Jordan, Fromme et al. 2001).

2.4.2. Cyclic electron flow

Cyclic electron flow (CEF) was first described over 50 years ago (Arnon, Allen et al.

1954). It occurs around PSI to promote the synthesis of ATP, and has an important function

in the avoidance of over-reduction on the PSI acceptor side (Munekage, Hashimoto et al.

2004). In CEF, electrons from the reducing side of PSI are re-injected into the photosynthetic

electron transport chain either at the level of the plastoquinone pool or at the stromal side

of the Cyt b6f (Joliot and Joliot 2006; Alric 2010). In green algae and higher plants, two

parallel cyclic electron transport pathways occur: NADPH dehydrogenase (NDH)-dependent

and Fd-dependent (proton gradient regulation5 (PGR5)-related) pathways (Okegawa, Long

et al. 2007; Shikanai 2007; Peltier, Tolleter et al. 2010).

The membrane bound PGR5 protein seems to have a role in electron transport from FNR to

the Cyt b6f (Munekage, Hojo et al. 2002; DalCorso, Pesaresi et al. 2008). Tight binding of

FNR to the Cyt b6f would provide a possible Fd-binding site for ferredoxin, and a pathway

for electron flow from the acceptor side of PSI to plastoquinone via Fd, FNR, and heme x in

the Cyt b6f. From plastoquinone, electrons can follow the normal pathway (via Cyt f and PC)

to P700 (Munekage, Hojo et al. 2002; Okegawa, Long et al. 2007). In the NDH-dependent

pathway, NDH mediates NADPH oxidation and plastoquinone pool reduction in similar way

as complex I does in mitochondria (Yamori, Sakata et al. 2011).

3. Photosystem I and Photosystem II complexes and their organization

Cyanobacteria, green algae and higher plants occupy a variety of ecological niches in

which they are exposed to changes in light intensities and qualities (Nelson and Ben-Shem

2005). To cope with the light changes, photosynthetic organisms must modulate their light-

harvesting machinery (Ben-Shem, Frolow et al. 2004). During the course of evolution,

Page 18: University of Groningen From Photosystem I to Photosystem

17

photosynthetic complexes acquired extrinsic and/or intrinsic peripheral light-harvesting

antenna complexes to maximize light harvesting and protect against high light damage

(Blankenship and Hartman 1998). PSI and PSII core complexes, which contains all the

cofactors of the electron transport chain, remain highly conserved in cyanobacteria, green

algae and higher plants (Nield, Kruse et al. 2000; Nield, Orlova et al. 2000; Jordan, Fromme

et al. 2001; Ben-Shem, Frolow et al. 2003; Buchel and Kuhlbrandt 2005; Umena, Kawakami

et al. 2011). In contrast, the structure, types of pigments, and organization of the antenna

complexes around reaction centers vary in different photosynthetic organisms (Amunts and

Nelson 2008; Blankenship 2010).

The proteins composing the PSI and PSII core complexes are encoded by the nuclear and

chloroplastic genomes and are referred to as Psa and Psb subunits, respectively (Green and

Durnford 1996). The genes encoding for the antenna of PSI and PSII from higher plants and

green algae, known as the Lhc multigene family, are nuclear encoded and have a range of

copy number. The Lhc proteins show structural homology, having three transmembrane

α-helices and coordinating Chl a, Chl b, and various carotenoid molecules (Jansson 1999).

3.1. Photosystem I

PSI of green algae and higher plants is a large protein complex that consists of two

functional units: the PSI core complex and the peripheral light harvesting antenna complex

(LHCI), which together form the PSI-LHCI supercomplex (Ben-Shem, Frolow et al. 2003).

In cyanobacteria, PSI core forms a trimer (Boekema, Dekker et al. 1987), while in green

algae and higher plants the PSI complex is monomeric (Lam, Oritz et al. 1984). Moreover, PSI

from green algae and higher plants is complemented by an asymmetric, heterogeneous, and

species specific set of light harvesting antenna complexes, which are completely different

from the phycobilisomes, which serve as a peripheral antenna in cyanobacteria

(Melkozernov, Barber et al. 2006; Busch and Hippler 2011).

3.1.1. PS I core complex

The central part of the core complex is formed by the two large transmembrane

protein subunits: PsaA and PsaB, comprising 22 transmembrane helices. These two subunits

bind the majority of the 100 Chl a that are associated with the core. This heterodimer is

surrounded by several small Psa subunits (Amunts and Nelson 2009). The functions of these

small PSI core subunits are: (1) binding or stabilizing electron transport cofactors such as FeS

Page 19: University of Groningen From Photosystem I to Photosystem

18

clusters FA and FB, (2) light harvesting and energy transfer, (3) docking and assembly of

trimerizing subunits, (4) photoprotection (Shi and Schroder 2004; Amunts and Nelson 2009).

The detailed function of each PSI subunit is summarized in Table 1.

In C. reinhardtii there are 14 Psa subunits (PsaA – PsaL, PsaN, PsaO) (Jensen, Bassi et al.

2007; Busch and Hippler 2011). The PSI core complexes of cyanobacteria and higher plants

show structural homology, however there are some differences. PsaX and PsaM subunits

are not present in photosynthetic eukaryotes (Jordan, Fromme et al. 2001; Ben-Shem,

Frolow et al. 2003), and four additional Psa subunits (PsaG, PsaH, PsaN, PsaR) are found in

higher plant structures (Ben-Shem, Frolow et al. 2003; Amunts, Toporik et al. 2010).

In addition, PsaO is not detected in the crystal structure from higher plants (Amunts,

Toporik et al. 2010; Busch and Hippler 2011).

Table 1. The characteristics of PSI core subunits of green algae and higher plants (Jensen,

Bassi et al. 2007; Amunts, Toporik et al. 2010).

Protein Molecular mass (kDa)

Cofactors Function

PsaA 83.2 Chl a, β-carotene Light harvesting

PsaB 82.5 P700, A0, A1, FX Charge separation, electron transport

PsaC 8.9 FA, FB Electron transport, binding of ferredoxin

PsaD 17.9 Binding of ferredoxin and PsaC

PsaE 10.4 Binding of ferredoxin and FNR, involved in cyclic electron transport

PsaF 17.3 Binding of plastocyanin and Lhca ¼

PsaG 11 1 Chl a, 1-2 β-carotene Binding of Lhca 1/4, regulation of PSI

PsaH 10.4 1 Chl a Binding of LHCII, stabilization of PsaD

PsaI 4.1 Stabilization of PsaL

PsaJ 5 2 Chl a Stabilization of PsaF

PsaK 8.5 2 Chl a Binding of Lhca 2/3

PsaL 18 3 Chl a Stabilization of PsaH and PsaO

PsaN 9.7 Docking of plastocyanin, stabilization of Lhca 2/3

PsaO 10.1 Binding of LHCII

PsaP* 10 Unknown

* not present in Chlamydomonas reinhardtii.

Page 20: University of Groningen From Photosystem I to Photosystem

19

3.1.2. LHCI

Both green alga and higher plants developed a highly sophisticated light harvesting

antenna system (LHC) that consists of one class of proteins (Jansson 1999). LHCI diverged

relatively early in the PSI evolution, and thus the stoichiometry and interaction with PSI

differ significantly between species. For example, C. reinhardtii does not possess the same

set of Lhca subunits found in higher plants (Durnford, Deane et al. 1999).

In higher plants, six genes (Lhca1–6) encode for the light-harvesting complex proteins

(Jansson 1999), Lhca5 and Lhca6 are only present in low amounts under normal growth

conditions (Ganeteg, Klimmek et al. 2004; Klimmek, Sjodin et al. 2006). The amino acid

sequences suggest that the structure of LHCI and LHCII is quite similar. Lhca1–4 are

composed of three transmembrane helices and they coordinate around 13-14 chlorophyll

molecules (Amunts, Toporik et al. 2010). In contrast to higher plants, C. reinhardtii has nine

genes coding Lhca proteins (Lhca1-Lhca9) which are all expressed ((Takahashi, Yasui et al.

2004; Drop, Webber-Birungi et al. 2011), chapter 2 of this thesis). A special characteristic of

PSI light harvesting antenna complexes, as compared to Lhcb in PSII, is the presence of the

so-called red chlorophylls. These “red forms” absorb light at wavelengths above 700 nm,

implying an uphill energy transfer to the P700 in PSI reaction center (Gobets and van

Grondelle 2001; Jennings, Zucchelli et al. 2003). In A. thaliana, Lhca4 as well as Lhca3

harbors the most red-shifted Chl which emits at 735 nm and 728 nm, respectively

(Castelletti, Morosinotto et al. 2003; Morosinotto, Mozzo et al. 2005; Wientjes and Croce

2011). The functional properties of the nine Lhcas of C. reinhardtii were studied previously

(Mozzo, Mantelli et al. 2010). C.r. Lhcas are divided into three subclasses: “blue Lhca”

(Lhca1, Lhca3 and Lhca7) with emission maxima at 682.5–683.5 nm, “intermediate Lhca”

(Lhca5, Lhca6 and Lhca8) with maxima between 694.5 and 697.5, and “red Lhca” (Lhca2,

Lhca4 and Lhca9) with maxima between 707 and 715 nm (Mozzo, Mantelli et al. 2010).

3.1.3. PSI-LHCI

The crystallization of the PSI-LHCI complex from pea and the subsequent

determination of X-ray structure (Ben-Shem, Frolow et al. 2003) was a landmark

achievement in PSI research (Jensen, Bassi et al. 2007). The most recent structure of higher

plant PSI-LHCI solved at 3.3 Å contains 18 protein subunits, 173 chlorophylls,

2 phylloquinones, 3 Fe4S4 clusters and 15 carotenoids (Fig. 4A) (Amunts, Toporik et al. 2010).

Page 21: University of Groningen From Photosystem I to Photosystem

20

A view from the stroma of the higher plant PSI complex reveals that the reaction centre and

LHCI form two distinct moieties, with a deep cleft between them (Fig. 4AB). The four

antenna proteins assemble into two dimers that are arranged in a halfmoon-shaped belt

and dock at the PsaF side of PSI core (Ben-Shem, Frolow et al. 2003). These two LHCI dimers

were identified as Lhca1/Lhca4 and Lhca2/Lhca3 and together correspond to a mass of 160

kDa. The whole PSI-LHCI has a mass of approximately 600 kDa (Amunts, Toporik et al. 2010).

The association of LHCI in PSI complex is more stable when compared with the association

of LHCII in the PSII supercomplex. However, this tighter association makes the PSI complex

less flexible (Morosinotto, Ballottari et al. 2005). The study of Lhcas mutants showed that

the binding of Lhca1/4 and Lhca2/3 dimers to the core is not interdependent. Moreover,

Lhca2 and Lhca4 can be associated with the PSI core even in the absence of their “dimeric

partners”, indicating that the docking sites for the individual subunits are highly specific

(Wientjes, Oostergetel et al. 2009).

Figure 4. The overall structure of higher plant PSI-LHCI. A: Structural model at 3.3 Å resolution (Amunts, Toporik et al. 2010). View from the stroma. B: The side view of the PSI structure embedded in thylakoid membranes (Nelson and Ben-Shem 2004).

A high-resolution structure of the PSI–LHCI complex from C. reinhardtii is not yet available.

Based on EM analysis and biochemical data it was concluded that PSI-LHCI is larger than that

of higher plants. Several models of C.r.PSI-LHCI, which contains between 6 and 14 Lhca

subunit, have been proposed (Bassi, Soen et al. 1992; Germano, Yakushevska et al. 2002;

Kargul, Nield et al. 2003; Takahashi, Yasui et al. 2004; Kargul, Turkina et al. 2005; Stauber,

Busch et al. 2009). The LHCI proteins were suggested to bind at the PsaF/J side of the

complex in one row (at the same positions as the four LHCI in higher plants) or two rows,

including the other side of the complex in the position at which in higher plants trimeric

Page 22: University of Groningen From Photosystem I to Photosystem

21

LHCII is bound in state 2 (Germano, Yakushevska et al. 2002; Kargul, Nield et al. 2003;

Kargul, Turkina et al. 2005). The recently published ((Drop, Webber-Birungi et al. 2011),

chapter 2 of this thesis) 2D maps of C.r.PSI-LHCI, isolated after mild detergent treatment,

shows 9 Lhcas organized on one side of the core in a double half-ring.

3.2. Photosystem II

Photosystem II of green algae and plants is organized into large supercomplexes

composed of PSII core complex associated with variable amounts of peripheral antenna

complexes (Kouril, Dekker et al. 2012).

3.2.1. PS II core complex

The PSII core complex (Fig. 5AB), which normally exists as a dimer (Peter and

Thornber 1991), has a total molecular weight of 350 kDa and its organization is very similar

in cyanobacteria, green alga and higher plants (Nield, Kruse et al. 2000; Nield, Orlova et al.

2000; Hankamer, Morris et al. 2001). The crystal structures of several cyanobacteria PSII

core complexes have been obtained at high resolution (Zouni, Witt et al. 2001; Kamiya and

Shen 2003; Ferreira, Iverson et al. 2004; Guskov, Kern et al. 2009; Umena, Kawakami et al.

2011).

Figure 5. Overall structure of PSII core complex. A: The PSII dimer from T. vulcanus at a resolution of 1.9Å. View from the direction perpendicular to the membrane (Umena, Kawakami et al. 2011); B: A stromal side view of the structure of the cyanobacterial PSII dimer. The boundary between the monomeric RCs is indicated (dashed arrow) (Nelson and Yocum 2006).

The current PSII crystal structures obtained from cyanobacteria at a resolution of 1.9Å

reveal the PSII core complex is composed of 20 transmembrane subunits and their cofactors

including: 35 chlorophylls a, 2 pheophytins, 11 β-carotenes and 2 plastoquinones, as well as

the metal atoms of the Mn4CaO5 cluster (Umena, Kawakami et al. 2011). In the middle of

the PSII core is the PSII reaction center complex in which the charge separation and primary

Page 23: University of Groningen From Photosystem I to Photosystem

22

electron transfer reactions take place (P680) (Dekker and Van Grondelle 2000). It is

composed of PsbA (D1) and PsbD (D2) proteins that bind 6 chlorophylls a and 2 pheophytins

(Hankamer, Morris et al. 2001). In close proximity to D1/D2 heterodimer, there are two low

molecular weight proteins PsbE and PsbF: α- and β- (respectively) subunits of cytochrome

b559 (Cyt b559). Both subunits, which are also part of PSII RC complex, form a heterodimer

and contain histidines providing ligands for binding a heme as cofactor. These are essential

for PSII activity, as their deletion results in the loss of PSII function and stability (Swiatek,

Regel et al. 2003). The RC complex is surrounded by PsbC (CP43) and PsbB (CP47), which

bind 13 Chl a and 16 Chl a, respectively (Umena, Kawakami et al. 2011). These proteins act

as an inner antenna and transfer excitation energy to the reaction center (Alfonso, Montoya

et al. 1994).

On the lumenal side of the PSII core complex, extrinsic proteins form the oxygen-evolving

complex (OEC), which is necessary for maintaining the water oxidation process (Bricker,

Roose et al. 2012). In green alga and higher plants OEC is composed of PsbO, PsbP, PsbQ

and PsbR (with molecular weight of 33 kDa, 23 kDa and 17 kDa respectively) (Nield, Kruse et

al. 2000; Nield, Orlova et al. 2000; Nield, Balsera et al. 2002; Ferreira, Iverson et al. 2004).

In cyanobacteria, the PsbO, PsbP and PsbQ homologue proteins are present along with PsbU

and PsbV (12 kDa and 15 kDa) (Thornton, Ohkawa et al. 2004; Roose, Kashino et al. 2007).

Additionally, a large number of low molecular weight (below 15kDa) intrinsic proteins are

associated with the PS II core (Shi and Schroder 2004; Roose, Wegener et al. 2007). Most of

them contain a single transmembrane helix, and their protein sequences are conserved

among photosynthetic organisms. Numerous biochemical, genetic and structural studies

have been used to probe the structure and function of these proteins within the

Photosystem. They are involved in stabilization, assembly or dimerization of the PSII

complex. However, the exact functions of most of them remain unknown (Shi and Schroder

2004; Shi, Hall et al. 2012). A summary of PSII core complex subunits is presented in Table 2.

Page 24: University of Groningen From Photosystem I to Photosystem

23

Table 2. The characteristics of PSII core subunits of green algae and higher plants

(Minagawa and Takahashi 2004; Shi and Schroder 2004; Muh, Renger et al. 2008; Shi, Hall et

al. 2012).

Protein Molecular

mass (kDa)

Function

PsbA (D1) 38.2 Involved in electron transport

PsbB (CP47) 56.1 Light harvesting

PsbC (CP43) 51.6 Light harvesting

PsbD (D2) 39.4 Primary reaction

PsbE 9.3 Involved in electron transport, PSII assembly and photoprotection

PsbF 4.9 Involved in electron transport, PSII assembly and photoprotection

PsbH 4.2 PSII stabilization and assembly, electron transport, photoprotection,

bicarbonate binding

PsbI 4.1 PSII dimerization/stabilization, maintenance of PSII structure and

function under high light

PsbJ 4.3 Assembly of water splitting complex, involved in electron transfer

within PSII

PsbK 4.4 Plastoquinone binding, maintaining PSII dimeric form

PsbL 3.7 Donor side electron transfer, assembly of PSII, maintaining PSII

dimeric form

PsbM 4.7 PSII dimer stabilization, reoxidation of PQ

PsbN 4 Unknown

PsbO 26.9 Mn clusters stabilization,

PsbP 23 Binding of Ca2+ and Cl

-

PsbQ 17 Unknown

PsbR 10 Associated with OEC, pH-dependent stabilizing protein for PSII,

docking protein for PSII extrinsic proteins

PsbS 21.7 Photoprotection,

PsbT 3.2 Recovery of photodamaged PSII, PSII dimerization/stabilization

PsbW 4.2 PSII dimerization/stabilization, photoprotection

PsbX 6.5 Binding or turnover of quinone molecules at QB site

PsbY 11 Unknown

PsbZ 6.5 Linker between LHCII and PSII core

Psb27 12.3 D1 C-terminal processing, Mn4Ca assembly, PSII assembly, PSII repair

Psb28 12.6 CP47 biogenesis, PSII assembly

Psb29 27.1 Regulation in photodamage and photorepair cycle

Psb30* 3.3 Stabilize PSII dimer, prevent Cytb559 from converting to low

potential form under high light

Psb32 22.4 Protection oxidative stress, repair PSII

* not present in Arabidopsis thaliana.

3.2.2. LHCII

The peripheral antenna complexes of PSII in green alga and higher plants belong to

the Lhc multigenic family. These complexes coordinate Chl a, Chl b and xanthophylls, in

different ratios (Jansson 1999). Two types of light harvesting antenna proteins associated

Page 25: University of Groningen From Photosystem I to Photosystem

24

with PSII can be distinguished: the major LHCII antenna complex is the most abundant and

in vivo is organized in heterotrimers, and the minor Lhcbs, which are present as monomers

(Dekker and Boekema 2005). The high-resolution structures of the major trimeric LHCII

complex (Kuhlbrandt, Wang et al. 1994; Liu, Yan et al. 2004; Standfuss, van Scheltinga et al.

2005) and the monomeric CP29 (Pan, Li et al. 2011) have been solved. A monomer of LHCII

binds 14 chlorophylls (8 Chls a and 6 Chls b) and 4 carotenoids (Liu, Yan et al. 2004).

In A. thaliana, LHCII proteins are encoded by Lhcb1 – Lhcb3 genes (Jansson 1999).

In contrast, in C. reinhardtii, the “major” LHCII proteins are encoded by genes, which based

on their sequence identity can be classified in four distinct types: type I (LhcbM3, M4, M6,

M8 and M9), type II (LhcbM5), type III (LhcbM2 and M7) and type IV (LhcbM1) (Teramoto,

Ono et al. 2001; Minagawa and Takahashi 2004). The minor monomeric Lhcb - CP29, CP26,

and CP24 - are encoded by Lhcb4, Lhcb5 and Lhcb6 genes respectively (Jansson 1999).

Whereas higher plants contain all three minor antennas, in C. reinhardtii only the first two

are present. Based on the sequences it is not possible to directly associate the different

LhcbM proteins of C.r. with Lhcb1 – Lhcb3 of A. thaliana. On the contrary, CP29 and CP26

are conserved (Teramoto, Ono et al. 2001; Minagawa and Takahashi 2004).

3.2.3. PSII-LHCII

PSII and LHCII form large supercomplexes in the thylakoid membranes consisting of

a dimeric core and a variable number of Lhcb subunits (Caffarri, Kouril et al. 2009).

The organization of PSII-LHCII complexes has been intensively studied by electron

microscopy and single particle analysis after mild solubilization of the membranes

(Boekema, van Roon et al. 1999; Boekema, van Roon et al. 1999; Nield, Kruse et al. 2000;

Nield, Orlova et al. 2000; Yakushevska, Jensen et al. 2001; Yakushevska, Keegstra et al. 2003;

Caffarri, Kouril et al. 2009; Tokutsu, Kato et al. 2012). The major LHCII can be distinguished

in three different types based on their strong (S), moderate (M) or loose (L) association with

the PSII core complex (Boekema, van Roon et al. 1999). The association of the peripheral

LHCII antenna to PSII core is extremely fragile and PSII-LHCII supercomplexes can be easily

disassembled (Caffarri, Kouril et al. 2009). This has been shown in studies where different

detergents in various concentrations resulted in a variation of the number of LHCII

associated with the PSII core complex (van Roon, van Breemen et al. 2000). The positions of

the S- and M- trimers are well defined in the supercomplex of higher plants. A dimeric core

Page 26: University of Groningen From Photosystem I to Photosystem

25

(C2) can associate with two LHCII-S trimers forming C2S2 supercomplex, and a further two

M-trimers are bound extending the supercomplex to C2S2M2 (Boekema, van Roon et al.

1999; Caffarri, Kouril et al. 2009). The position of loosely bound trimers in PSII-LHCII

supercomplex of higher plants is unclear. Although LHCII-L has been reported in PSII-LHCII

from spinach, in which PSII supercomplexes are organized as C2S2M1L1, C2S2M2L1, C2S2L1,

the presence of the most intact structure of PSII-LHCII with a C2S2M2L2 organization has not

been confirmed (Boekema, van Roon et al. 1999). A very mild solubilization of the thylakoid

membranes with low concentration of α–DM was successfully used to investigate structures

of photosynthetic complexes in A. thaliana. In this study, the C2S2M2 supercomplex was

the PSII-LHCII with the largest area, and the L-trimer association was not observed (Caffarri,

Kouril et al. 2009).

A comparison of the PSII structures from C.r. and spinach based on their 3D map was shown

by Nield and co-workers (Nield, Kruse et al. 2000; Nield, Orlova et al. 2000). Both models are

very similar in shape and size. The C.r. PSII dimeric core complex is associated with

LHCII/CP29/CP26 proteins that are accommodated in a structure similar to higher plants,

suggesting that the C2S2 is the basic unit of PSII supercomplexes in eukaryotic organisms

(Nield, Kruse et al. 2000). The lack of both M- or L- trimers or their relatively weak binding to

the core has been linked to the absence of CP24, a minor monomeric LHCII protein specific

to higher plants that was suggested to serve as a linker between PSII core subunits and

M- and L- trimers in higher plants (Kovacs, Damkjaer et al. 2006; de Bianchi, Dall'Osto et al.

2008). However, a very recent study shows that the PSII supercomplex of C.r. actually

contains three LHCII trimers per RC ((Tokutsu, Kato et al. 2012; Drop, Webber-Birungi et al.

2014) and chapter 3 of this thesis).

4. Photoacclimation and photoprotection

In nature, photosynthetic organisms are constantly exposed to changes in light

quality and quantity, and must modulate the operation of the photosynthetic complexes in

response to these changes (Lemeille and Rochaix 2010). Depending on the timescale of

activation, these mechanisms are classified as short-term (direct effects; in the timescale of

seconds or minutes) and long-term (at the level of gene expression; in the timescale of

hours or days) responses (Scheibe, Backhausen et al. 2005). The short-term responses may

reflect changes in protonation, phosphorylation, and/or the association of various pigment

Page 27: University of Groningen From Photosystem I to Photosystem

26

and protein components of the photosynthetic complexes. Longer-term responses can

result in changes in subunit stoichiometries, pigment composition, and the insertion of

novel proteins into individual complexes (Grossman, Karpowicz et al. 2010).

4.1. High light conditions

Light is necessary for photosynthesis, however the excess of illumination leads to

saturation of linear electron transport chain and damage of photosynthetic complexes.

Green algae and higher plants have developed a variety of mechanisms to cope with high

light stress (Niyogi 1999).

4.1.1. High light acclimation

Long-lasting high light illumination triggers acclimation responses in green algae and higher

plants acclimation responses. This process in plants includes adjustment of reaction centre

stoichiometry and modulation of the light harvesting antenna size (Chow, Melis et al. 1990;

Bailey, Walters et al. 2001; Ballottari, Dall'Osto et al. 2007; Wientjes, van Amerongen et al.

2013).

The acclimation mechanisms in green algae differ between species. To adapt to high light

Dunaliella salina reduces the antenna size of both photosystems as well as the PSI/PSII ratio

(Smith, Morrissey et al. 1990). In contrast, Dunaliella tertiolecta changes only the PSI/PSII

stoichiometry by decreasing the number of PSI reaction centers per cell (Falkowski and

Owens 1980). The high light acclimation mechanisms in C. reinhardtii remain obscure.

Durnford and co-workers reported a decrease in the transcription and translation of Lhc

proteins in response to high light (McKim and Durnford 2006). In contrast, the recent work

of Bonente and co-workers on C. reinhardtii acclimated to high light showed that the

antenna size of both PSI and PSII was identical in all conditions, and that acclimation of

C. reinhardtii to high light is regulated only by the PSI/PSII ratio (Bonente, Pippa et al. 2012).

4.1.2. Short term responses

On a timescale of seconds to minutes, photosynthetic organisms protect themselves

from rapid increases in light intensity by the process known as nonphotochemical quenching

(NPQ) (Niyogi, Li et al. 2005). The NPQ mechanism refers to quenching of chlorophyll a

fluorescence, which is induced under steady-state illumination. The NPQ is be divided into

Page 28: University of Groningen From Photosystem I to Photosystem

27

three different components, which are: (1) ΔpH-dependent quenching (qE), (2) state

transition (qT), and (3) photoinhibitory quenching (qI) (Muller, Li et al. 2001). qE is the

fastest component of NPQ in which the excess of absorbed light energy is dissipated as heat

(Horton, Ruban et al. 1996). This reduces the possibility for formation of reactive oxygen

species (ROS) and thus photo oxidation (Scheibe, Backhausen et al. 2005). ROS can damage

not only photosynthetic complexes, but also other proteins, membranes and DNA, and can

lead to cell death (Fischer, Wiesendanger et al. 2006). qE is controlled by the trans-thylakoid

pH gradient. In higher plants, the development of this process is correlated with the

xanthophyll cycle (Gilmore, Hazlett et al. 1995; Horton, Ruban et al. 1996). In this process,

low lumenal pH activates the violaxanthin de-epoxidase (VDE), which converts violaxanthin

to antheraxanthin and then to zeaxanthin (Fig. 6) (Demmig, Winter et al. 1987; Niyogi,

Grossman et al. 1998).

Figure 6. The xanthophyll cycle. The de-epoxidation from violaxanthin to zeaxanthin by VDE occurs on the lumen side and epoxidation by zeaxanthin epoxidase (ZE) on the stroma side. Light-induced acidification of the lumen below pH 6 induces the binding of VDE to the thylakoid membrane. The epoxidase is bound to the thyakoid on the stroma side and catalyzes epoxidation at pH 7.0.

In higher plants, low lumenal pH also leads to protonation of PsbS – a Lhc-homologous PSII

subunit which has been shown to be essential for qE (Li, Bjorkman et al. 2000).

The hypothesis is that the formation of zeaxanthin and the activation of PsbS induces

changes in the membrane organization and in the LHCII conformation that promote thermal

dissipation (Havaux and Niyogi 1999; Li, Gilmore et al. 2004; Niyogi, Li et al. 2005; Ruban,

Berera et al. 2007).

Page 29: University of Groningen From Photosystem I to Photosystem

28

Although the PsbS gene is present in C. reinhardtii genome, it was not found to be expressed

in various experimental conditions (Anwaruzzaman, Chin et al. 2004; Bonente, Passarini et

al. 2008). In addition, it was shown that C.r. mutant deficient in violaxanthin deepoxidase

activity still exhibit qE, suggesting the xanthophyll cycle is not required for photoprotection

in C. reinhardtii (Niyogi, Bjorkman et al. 1997; Niyogi, Bjorkman et al. 1997). These results

suggest a different qE mechanism in this green alga. Indeed, it was shown that a mutant

deficient in LhcSR have a strong reduction of qE compared to wildtype, indicating that the

activation of qE in C. reinhardtii is dependent on the accumulation of LhcSR, which occurs

only in high light (Peers, Truong et al. 2009). LhcSR is a member of the Lhc family, and in

C. reinhardtii is encoded by lhcsr1 and lhcsr3 genes (Peers, Truong et al. 2009; Bonente,

Ballottari et al. 2011). LhcSR is found in algae and mosses (Elrad and Grossman, 2004;

Nymark et al., 2009) and in contrast to PsbS, it binds pigments – Chl a, Chl b, lutein and

violaxanthin (Bonente, Ballottari et al. 2011). Recently, it was suggests that the association

of LhcSR with PSII-LHCII under high light condition is essential for qE quenching in

C. reinhardtii (Tokutsu and Minagawa 2013). During high light illumination, LhcSR associates

with PSII-LHCII to form the PSII-LHCII-LhcSR supercomplex. In the next step, protonation of

LhcSR modifies the antenna conformation within the PSII supercomplex to form a quenching

center (Tokutsu and Minagawa 2013).

4.2. State transitions

Although both PSI and PSII contain Chls and Cars, each of the photosystems has

a distinct pigment composition with distinct absorption characteristics. LHCII contains more

chlorophyll b than LHCI, but LHCI contains “red-shifted” chlorophylls able to absorb light at

wavelengths longer than 700 nm (Wientjes and Croce, 2011). As the result, PSII and PSI

complexes more effectively absorb red light (around 650 nm) and far-red light (around 700

nm) respectively (Veeranjaneyulu and Leblanc 1994). In the linear electron transport, PSI

and PSII must operate at the same rate (Amunts and Nelson 2008), but when the light

quality and quantity fluctuate with time, conditions that favor the excitation of one complex

over the other may occur. This could create an imbalance of energy distribution between

PSII and PSI, which would potentially decrease the overall efficiency of photosynthesis (Allen

1992; Iwai, Takahashi et al. 2008). Higher plants and green algae compensate for this via

a short-term adaptation mechanism known as state transitions (Bonaventura and Myers

Page 30: University of Groningen From Photosystem I to Photosystem

29

1969; Murata 1969). In this process the peripheral antenna systems of PSI and PSII are

adjusted to balance the effective absorption cross-sections of the two photosystems (Allen

and Forsberg 2001). In state 1 a mobile part of LHCII is associated with PSII. In state 2, LHCII

detaches from PSII and migrates from the PSII-rich grana stacks of the thylakoid membranes

toward the unstacked membranes of the stroma, where PSI is located (Andersson, Akerlund

et al. 1982; Vallon, Bulte et al. 1991; Delosme, Olive et al. 1996; Iwai, Takahashi et al. 2008).

4.2.1. Mechanism of state transitions

The current model of state transitions assumes that phosphorylation of LHCII is

required for the transition from state 1 to state 2 (Rochaix 2007). LHCII is phosphorylated by

a thylakoid protein kinase – stt7 in C. reinhardtii and its homologue stn7 in A. thaliana

(Depege, Bellafiore et al. 2003; Bellafiore, Barneche et al. 2005; Bonardi, Pesaresi et al.

2005). The activity of the LHCII kinase is regulated by co-operative redox control, both via

PQ and Cyt b6f (Vener, Van Kan et al. 1995). When PSII is overexcited relative to PSI,

the plastoquinone pool becomes over-reduced. The docking of plastoquinol to the Qo of

Cyt b6f complex activates the LHCII kinase by the reduction of thiol groups (Vener, VanKan

et al. 1997; Zito, Finazzi et al. 1999). This event triggers the phosphorylation of LHCII, its

release from PSII and migration to PSI (Allen 1992). Reversely, under PSI-favoring light

conditions, the PQ pool is oxidized, leading to the de-activation of the LHCII-specific kinases

and to dephosphorylation of LHCII by the TAP38/PPH1 phosphatase. Dephosphorylated

LHCII detaches from PSI and reassociates with PSII (Bennett 1980; Pribil, Pesaresi et al. 2010;

Shapiguzov, Ingelsson et al. 2010).

However, the recent study of Wientjes and co-workers on the structural and functional

organization of A.t PSI and PSII under different light intensities showed a new view on light

acclimation and state transition process in higher plants. In this study, it was shown that

a portion of “extra” (non-PSII supercomplex) LHCII is part of the PSI antenna complex under

all growth conditions (Wientjes, van Amerongen et al. 2013).

4.2.2. Structural changes during the state transitions process

It is a widely accepted view that during state 1 to state 2 transition as much as 80%

of the LHC antenna is mobile in C. reindhardtii, whereas in higher plants only 15–20% of

LHCII dissociates from PSII and attaches to PSI (Allen 1992; Delosme, Olive et al. 1996).

However, recent work using time-resolved fluorescence measurements to study the

Page 31: University of Groningen From Photosystem I to Photosystem

30

changes in the PSI and PSII antenna size in state 1 and state 2 for C. reinhardtii

demonstrated that although LHCII does indeed detach from PSII in state 2, only a small

fraction actually attaches to PSI (Unlu, Drop et al. 2014).

State transitions also involve remodeling of the thylakoid membranes (Lemeille and Rochaix

2010). Analysis of the thylakoid membranes of Arabidopsis thaliana during state transitions

by atomic force microscopy, scanning and transmission electron microscopy and confocal

imaging allow the proposal of a model in which the reorganization of the membranes is

mediated by fusion and fission events at the interface between the granal and stromal

lamellar domains of the thylakoid membranes (Chuartzman, Nevo et al. 2008).

The model of thylakoid membranes remodeling during state transitions in C. reinhardtii was

proposed by (Iwai, Takahashi et al. 2008). In the first step, the phosphorylation of LHCII

leads to division of the PSII megacomplex into PSII supercomplexes that next leads to

unstacking of the thylakoid membranes. In the next step, the phosphorylation of CP26 and

CP29 as well as PSII core complex subunits initiates the undocking of LHCII from the PSII

core. The phosphorylated free LHCII would then re-associate with the PSI (Iwai, Takahashi et

al. 2008).

4.2.3. PSI-LHCI-LHCII supercomplex

State transitions studies in higher plant indicated that several peripheral PSI core

subunits are involved in the interactions between LHCII and PSI. Analysis of A. thaliana PSI

mutants lacking either PsaH, PsaI, PsaL or PsaO showed strong reduction of state transitions,

which suggest that these subunits possibly form a docking site for LHCII (Lunde, Jensen et al.

2000; Jensen, Haldrup et al. 2004; Zhang and Scheller 2004). The analysis of X-ray structure

of higher plant PSI (Ben-Shem, Frolow et al. 2003), showed that the PsaH protein is located

at an exposed hydrophobic surface of PSI and binds a chlorophyll molecule which could then

participate in the energy transfer from LHCII to the PSI complex (Amunts and Nelson 2008).

The 16 Å resolution EM data of single particle analysis of PSI-LHCI in state 2 from A.t.

reported a large density along the side of PsaH/PsaL/PsaA/PsaK, which was assigned to the

LHCII trimer (Kouril, Zygadlo et al. 2005). So far, there is no evidence that in higher plants

monomeric antenna also participate in state transitions.

In C. reindhardtii it was shown that two minor Lhcb proteins (CP26 and CP29) and one

major LHCII protein LhcbM5 were bound to PSI-LHCI complex isolated in state 2. It has been

Page 32: University of Groningen From Photosystem I to Photosystem

31

hypothesized that these monomeric antennas may act as linker proteins between the LHCII

and the PSI core (Takahashi, Iwai et al. 2006). A further study demonstrated that under state

2 conditions, CP29 is a docking protein for LHCII and is crucial for the association of the

mobile fraction of LHCII to PSI (Tokutsu, Iwai et al. 2009). However, single particle analysis of

C.r.PSI-LHCI isolated in state 2 showed only two monomeric Lhcbs, which were suggested to

be CP29, associated with the C.r.PSI on both sides of the PsaH subunit (Kargul, Turkina et al.

2005).

Recently, a biochemical study identified a supercomplex in C. reinhardtii composed of the

PSI-LHCI supercomplex with LHCIIs, Cyt b6f, Fd–NADPH oxidoreductase (FNR) and the

integral membrane protein PGRL1. It was shown the formation of this super-supercomplex

serves as the molecular switch for cyclic electron flow (CEF) (Iwai, Takizawa et al. 2010), but

no structural data was reported.

Page 33: University of Groningen From Photosystem I to Photosystem

32

References

Alfonso, M., G. Montoya, et al. (1994). "Core antenna complexes, CP43 and CP47, of higher plant photosystem II. Spectral properties, pigment stoichiometry, and amino acid composition." Biochemistry 33(34): 10494-10500.

Allen, J. F. (1992). "Protein-Phosphorylation in Regulation of Photosynthesis." Biochimica Et Biophysica Acta 1098(3): 275-335.

Allen, J. F., W. B. de Paula, et al. (2011). "A structural phylogenetic map for chloroplast photosynthesis." Trends in Plant Science 16(12): 645-655.

Allen, J. F. and J. Forsberg (2001). "Molecular recognition in thylakoid structure and function." Trends in Plant Science 6(7): 317-326.

Alric, J. (2010). "Cyclic electron flow around photosystem I in unicellular green algae." Photosynthesis Research 106(1-2): 47-56.

Amunts, A. and N. Nelson (2008). "Functional organization of a plant Photosystem I: evolution of a highly efficient photochemical machine." Plant Physiol Biochem 46(3): 228-237.

Amunts, A. and N. Nelson (2009). "Plant photosystem I design in the light of evolution." Structure 17(5): 637-650.

Amunts, A., H. Toporik, et al. (2010). "Structure determination and improved model of plant photosystem I." Journal of Biological Chemistry 285(5): 3478-3486.

Andersson, B., H. E. Akerlund, et al. (1982). "Differential Phosphorylation of the Light-Harvesting Chlorophyll Protein Complex in Appressed and Non-Appressed Regions of the Thylakoid Membrane." Febs Letters 149(2): 181-185.

Anwaruzzaman, M., B. L. Chin, et al. (2004). "Genomic analysis of mutants affecting xanthophyll biosynthesis and regulation of photosynthetic light harvesting in Chlamydomonas reinhardtii." Photosynthesis Research 82(3): 265-276.

Arnon, D. I. (1971). "The light reactions of photosynthesis." Proc Natl Acad Sci U S A 68(11): 2883-2892.

Arnon, D. I., M. B. Allen, et al. (1954). "Photosynthesis by isolated chloroplasts." Nature 174(4426): 394-396.

Bailey, S., R. G. Walters, et al. (2001). "Acclimation of Arabidopsis thaliana to the light environment: the existence of separate low light and high light responses." Planta 213(5): 794-801.

Ballottari, M., L. Dall'Osto, et al. (2007). "Contrasting behavior of higher plant photosystem I and II antenna systems during acclimation." Journal of Biological Chemistry 282(12): 8947-8958.

Barber, J. (2002). "Photosystem II: a multisubunit membrane protein that oxidises water." Curr Opin Struct Biol 12(4): 523-530.

Barber, J. (2004). "Engine of life and big bang of evolution: a personal perspective." Photosynthesis Research 80(1-3): 137-155.

Bassi, R., S. Y. Soen, et al. (1992). "Characterization of chlorophyll a/b proteins of photosystem I from Chlamydomonas reinhardtii." Journal of Biological Chemistry 267(36): 25714-25721.

Beja, O., E. N. Spudich, et al. (2001). "Proteorhodopsin phototrophy in the ocean." Nature 411(6839): 786-789.

Bellafiore, S., F. Barneche, et al. (2005). "State transitions and light adaptation require chloroplast thylakoid protein kinase STN7." Nature 433(7028): 892-895.

Ben-Shem, A., F. Frolow, et al. (2003). "Crystal structure of plant photosystem I." Nature 426(6967): 630-635.

Ben-Shem, A., F. Frolow, et al. (2004). "Light-harvesting features revealed by the structure of plant photosystem I." Photosynthesis Research 81(3): 239-250.

Bennett, J. (1980). "Chloroplast Phosphoproteins - Evidence for a Thylakoid-Bound Phosphoprotein Phosphatase." European Journal of Biochemistry 104(1): 85-89.

Blankenship, R. E. (1992). "Origin and early evolution of photosynthesis." Photosynthesis Research 33: 91-111.

Blankenship, R. E. (2010). "Early evolution of photosynthesis." Plant Physiology 154(2): 434-438.

Page 34: University of Groningen From Photosystem I to Photosystem

33

Blankenship, R. E. and H. Hartman (1998). "The origin and evolution of oxygenic photosynthesis." Trends in Biochemical Sciences 23(3): 94-97.

Boekema, E. J., J. P. Dekker, et al. (1987). "Evidence for a Trimeric Organization of the Photosystem-I Complex from the Thermophilic Cyanobacterium Synechococcus Sp." Febs Letters 217(2): 283-286.

Boekema, E. J., H. van Roon, et al. (1999). "Multiple types of association of photosystem II and its light-harvesting antenna in partially solubilized photosystem II membranes." Biochemistry 38(8): 2233-2239.

Boekema, E. J., H. van Roon, et al. (1999). "Supramolecular organization of photosystem II and its light-harvesting antenna in partially solubilized photosystem II membranes." European Journal of Biochemistry 266(2): 444-452.

Bonardi, V., P. Pesaresi, et al. (2005). "Photosystem II core phosphorylation and photosynthetic acclimation require two different protein kinases." Nature 437(7062): 1179-1182.

Bonaventura, C. and J. Myers (1969). "Fluorescence and oxygen evolution from Chlorella pyrenoidosa." Biochimica Et Biophysica Acta 189(3): 366-383.

Bonente, G., M. Ballottari, et al. (2011). "Analysis of LhcSR3, a protein essential for feedback de-excitation in the green alga Chlamydomonas reinhardtii." Plos Biology 9(1): e1000577.

Bonente, G., F. Passarini, et al. (2008). "The occurrence of the psbS gene product in Chlamydomonas reinhardtii and in other photosynthetic organisms and its correlation with energy quenching." Photochemistry and Photobiology 84(6): 1359-1370.

Bonente, G., S. Pippa, et al. (2012). "Acclimation of Chlamydomonas reinhardtii to different growth irradiances." Journal of Biological Chemistry 287(8): 5833-5847.

Bricker, T. M., J. L. Roose, et al. (2012). "The extrinsic proteins of Photosystem II." Biochimica Et Biophysica Acta 1817(1): 121-142.

Buchel, C. and W. Kuhlbrandt (2005). "Structural differences in the inner part of photosystem II between higher plants and cyanobacteria." Photosynthesis Research 85(1): 3-13.

Busch, A. and M. Hippler (2011). "The structure and function of eukaryotic photosystem I." Biochimica Et Biophysica Acta 1807(8): 864-877.

Caffarri, S., R. Kouril, et al. (2009). "Functional architecture of higher plant photosystem II supercomplexes." Embo Journal 28(19): 3052-3063.

Castelletti, S., T. Morosinotto, et al. (2003). "Recombinant Lhca2 and Lhca3 subunits of the photosystem I antenna system." Biochemistry 42(14): 4226-4234.

Chow, W. S., A. Melis, et al. (1990). "Adjustments of photosystem stoichiometry in chloroplasts improve the quantum efficiency of photosynthesis." Proc Natl Acad Sci U S A 87(19): 7502-7506.

Chuartzman, S. G., R. Nevo, et al. (2008). "Thylakoid membrane remodeling during state transitions in Arabidopsis." Plant Cell 20(4): 1029-1039.

DalCorso, G., P. Pesaresi, et al. (2008). "A complex containing PGRL1 and PGR5 is involved in the switch between linear and cyclic electron flow in Arabidopsis." Cell 132(2): 273-285.

Davies, J. P. and A. R. Grossman (1998). "The use of Chlamydomonas (Chlorophyta : Volvocales) as a model algal system for genome studies and the elucidation of photosynthetic processes." Journal of Phycology 34(6): 907-917.

de Bianchi, S., L. Dall'Osto, et al. (2008). "Minor antenna proteins CP24 and CP26 affect the interactions between photosystem II Subunits and the electron transport rate in grana membranes of Arabidopsis." Plant Cell 20(4): 1012-1028.

Debus, R. J. (1992). "The manganese and calcium ions of photosynthetic oxygen evolution." Biochimica Et Biophysica Acta 1102(3): 269-352.

Deisenhofer, J. and H. Michel (1991). "Structures of bacterial photosynthetic reaction centers." Annu Rev Cell Biol 7: 1-23.

Dekker, J. P. and E. J. Boekema (2005). "Supramolecular organization of thylakoid membrane proteins in green plants." Biochimica Et Biophysica Acta 1706(1-2): 12-39.

Page 35: University of Groningen From Photosystem I to Photosystem

34

Dekker, J. P. and R. Van Grondelle (2000). "Primary charge separation in Photosystem II." Photosynthesis Research 63(3): 195-208.

Delosme, R., J. Olive, et al. (1996). "Changes in light energy distribution upon state transitions: An in vivo photoacoustic study of the wild type and photosynthesis mutants from Chlamydomonas reinhardtii." Biochimica Et Biophysica Acta-Bioenergetics 1273(2): 150-158.

Demmig, B., K. Winter, et al. (1987). "Photoinhibition and zeaxanthin formation in intact leaves : a possible role of the xanthophyll cycle in the dissipation of excess light energy." Plant Physiology 84(2): 218-224.

Depege, N., S. Bellafiore, et al. (2003). "Role of chloroplast protein kinase Stt7 in LHCII phosphorylation and state transition in Chlamydomonas." Science 299(5612): 1572-1575.

Drop, B., M. Webber-Birungi, et al. (2011). "Photosystem I of Chlamydomonas reinhardtii contains nine light-harvesting complexes (Lhca) located on one side of the core." Journal of Biological Chemistry 286(52): 44878-44887.

Drop, B., M. Webber-Birungi, et al. (2014). "Light-harvesting complex II (LHCII) and its supramolecular organization in Chlamydomonas reinhardtii." Biochim Biophys Acta 1837(1): 63-72.

Durnford, D. G., J. A. Deane, et al. (1999). "A phylogenetic assessment of the eukaryotic light-harvesting antenna proteins, with implications for plastid evolution." J Mol Evol 48(1): 59-68.

Falkowski, P. G. and T. G. Owens (1980). "Light-Shade Adaptation : TWO STRATEGIES IN MARINE PHYTOPLANKTON." Plant Physiology 66(4): 592-595.

Ferreira, K. N., T. M. Iverson, et al. (2004). "Architecture of the photosynthetic oxygen-evolving center." Science 303(5665): 1831-1838.

Fischer, B. B., M. Wiesendanger, et al. (2006). "Growth condition-dependent sensitivity, photodamage and stress response of Chlamydomonas reinhardtii exposed to high light conditions." Plant and Cell Physiology 47(8): 1135-1145.

Ganeteg, U., F. Klimmek, et al. (2004). "Lhca5--an LHC-type protein associated with photosystem I." Plant Molecular Biology 54(5): 641-651.

Germano, M., A. E. Yakushevska, et al. (2002). "Supramolecular organization of photosystem I and light-harvesting complex I in Chlamydomonas reinhardtii." Febs Letters 525(1-3): 121-125.

Gilmore, A. M., T. L. Hazlett, et al. (1995). "Xanthophyll cycle-dependent quenching of photosystem II chlorophyll a fluorescence: formation of a quenching complex with a short fluorescence lifetime." Proc Natl Acad Sci U S A 92(6): 2273-2277.

Gobets, B. and R. van Grondelle (2001). "Energy transfer and trapping in photosystem I." Biochimica Et Biophysica Acta 1507(1-3): 80-99.

Green, B. R. and D. G. Durnford (1996). "The Chlorophyll-Carotenoid Proteins of Oxygenic Photosynthesis." Annu Rev Plant Physiol Plant Mol Biol 47: 685-714.

Grossman, A. R., S. J. Karpowicz, et al. (2010). "Phylogenomic analysis of the Chlamydomonas genome unmasks proteins potentially involved in photosynthetic function and regulation." Photosynthesis Research 106(1-2): 3-17.

Guskov, A., J. Kern, et al. (2009). "Cyanobacterial photosystem II at 2.9-angstrom resolution and the role of quinones, lipids, channels and chloride." Nat Struct Mol Biol 16(3): 334-342.

Hankamer, B., E. Morris, et al. (2001). "Three-dimensional structure of the photosystem II core dimer of higher plants determined by electron microscopy." J Struct Biol 135(3): 262-269.

Harris, E. H. (2001). "Chlamydomonas as a model organism." Annu Rev Plant Physiol Plant Mol Biol 52: 363-406.

Harris, E. H. (2008). The Chlamydomonas Sourcebook, Academic Press. Havaux, M. and K. K. Niyogi (1999). "The violaxanthin cycle protects plants from photooxidative

damage by more than one mechanism." Proc Natl Acad Sci U S A 96(15): 8762-8767. Horton, P., A. V. Ruban, et al. (1996). "Regulation of Light Harvesting in Green Plants." Annu Rev

Plant Physiol Plant Mol Biol 47: 655-684.

Page 36: University of Groningen From Photosystem I to Photosystem

35

Iwai, M., Y. Takahashi, et al. (2008). "Molecular remodeling of photosystem II during state transitions in Chlamydomonas reinhardtii." Plant Cell 20(8): 2177-2189.

Iwai, M., K. Takizawa, et al. (2010). "Isolation of the elusive supercomplex that drives cyclic electron flow in photosynthesis." Nature 464(7292): 1210-1213.

Jansson, S. (1999). "A guide to the Lhc genes and their relatives in Arabidopsis." Trends in Plant Science 4(6): 236-240.

Jennings, R. C., G. Zucchelli, et al. (2003). "The photochemical trapping rate from red spectral states in PSI-LHCI is determined by thermal activation of energy transfer to bulk chlorophylls." Biochimica Et Biophysica Acta 1557(1-3): 91-98.

Jensen, P. E., R. Bassi, et al. (2007). "Structure, function and regulation of plant photosystem I." Biochimica Et Biophysica Acta 1767(5): 335-352.

Jensen, P. E., A. Haldrup, et al. (2004). "The PSI-O subunit of plant photosystem I is involved in balancing the excitation pressure between the two photosystems." Journal of Biological Chemistry 279(23): 24212-24217.

Joliot, P. and A. Joliot (2006). "Cyclic electron flow in C3 plants." Biochimica Et Biophysica Acta 1757(5-6): 362-368.

Jordan, P., P. Fromme, et al. (2001). "Three-dimensional structure of cyanobacterial photosystem I at 2.5 A resolution." Nature 411(6840): 909-917.

Kamiya, N. and J. R. Shen (2003). "Crystal structure of oxygen-evolving photosystem II from Thermosynechococcus vulcanus at 3.7-angstrom resolution." Proceedings of the National Academy of Sciences of the United States of America 100(1): 98-103.

Kargul, J., J. Nield, et al. (2003). "Three-dimensional reconstruction of a light-harvesting complex I-photosystem I (LHCI-PSI) supercomplex from the green alga Chlamydomonas reinhardtii. Insights into light harvesting for PSI." Journal of Biological Chemistry 278(18): 16135-16141.

Kargul, J., M. V. Turkina, et al. (2005). "Light-harvesting complex II protein CP29 binds to photosystem I of Chlamydomonas reinhardtii under State 2 conditions." Febs Journal 272(18): 4797-4806.

Klimmek, F., A. Sjodin, et al. (2006). "Abundantly and rarely expressed Lhc protein genes exhibit distinct regulation patterns in plants." Plant Physiology 140(3): 793-804.

Kok, B., B. Forbush, et al. (1970). "Cooperation of charges in photosynthetic O2 evolution-I. A linear four step mechanism." Photochemistry and Photobiology 11(6): 457-475.

Kouril, R., J. P. Dekker, et al. (2012). "Supramolecular organization of photosystem II in green plants." Biochimica Et Biophysica Acta 1817(1): 2-12.

Kouril, R., A. Zygadlo, et al. (2005). "Structural characterization of a complex of photosystem I and light-harvesting complex II of Arabidopsis thaliana." Biochemistry 44(33): 10935-10940.

Kovacs, L., J. Damkjaer, et al. (2006). "Lack of the light-harvesting complex CP24 affects the structure and function of the grana membranes of higher plant chloroplasts." Plant Cell 18(11): 3106-3120.

Kuhlbrandt, W., D. N. Wang, et al. (1994). "Atomic Model of Plant Light-Harvesting Complex by Electron Crystallography." Nature 367(6464): 614-621.

Kurisu, G., H. Zhang, et al. (2003). "Structure of the cytochrome b6f complex of oxygenic photosynthesis: tuning the cavity." Science 302(5647): 1009-1014.

Lam, E., W. Oritz, et al. (1984). "Isolation and Characterization of a Light-Harvesting Chlorophyll a/b Protein Complex Associated with Photosystem I." Plant Physiology 74(3): 650-655.

Lemeille, S. and J. D. Rochaix (2010). "State transitions at the crossroad of thylakoid signalling pathways." Photosynthesis Research 106(1-2): 33-46.

Li, X. P., O. Bjorkman, et al. (2000). "A pigment-binding protein essential for regulation of photosynthetic light harvesting." Nature 403(6768): 391-395.

Li, X. P., A. M. Gilmore, et al. (2004). "Regulation of photosynthetic light harvesting involves intrathylakoid lumen pH sensing by the PsbS protein." Journal of Biological Chemistry 279(22): 22866-22874.

Page 37: University of Groningen From Photosystem I to Photosystem

36

Liu, Z. F., H. C. Yan, et al. (2004). "Crystal structure of spinach major light-harvesting complex at 2.72 angstrom resolution." Nature 428(6980): 287-292.

Lunde, C., P. E. Jensen, et al. (2000). "The PSI-H subunit of photosystem I is essential for state transitions in plant photosynthesis." Nature 408(6812): 613-615.

Mayfield, S. P. and S. E. Franklin (2005). "Expression of human antibodies in eukaryotic micro-algae." Vaccine 23(15): 1828-1832.

McKim, S. M. and D. G. Durnford (2006). "Translational regulation of light-harvesting complex expression during photoacclimation to high-light in Chlamydomonas reinhardtii." Plant Physiol Biochem 44(11-12): 857-865.

Melis, A., L. P. Zhang, et al. (2000). "Sustained photobiological hydrogen gas production upon reversible inactivation of oxygen evolution in the green alga Chlamydomonas reinhardtii." Plant Physiology 122(1): 127-135.

Melkozernov, A. N., J. Barber, et al. (2006). "Light harvesting in photosystem I supercomplexes." Biochemistry 45(2): 331-345.

Merchant, S. S., S. E. Prochnik, et al. (2007). "The Chlamydomonas genome reveals the evolution of key animal and plant functions." Science 318(5848): 245-250.

Minagawa, J. and Y. Takahashi (2004). "Structure, function and assembly of Photosystem II and its light-harvesting proteins." Photosynthesis Research 82(3): 241-263.

Mojzsis, S. J., G. Arrhenius, et al. (1996). "Evidence for life on Earth before 3,800 million years ago." Nature 384(6604): 55-59.

Morosinotto, T., M. Ballottari, et al. (2005). "The association of the antenna system to photosystem I in higher plants. Cooperative interactions stabilize the supramolecular complex and enhance red-shifted spectral forms." Journal of Biological Chemistry 280(35): 31050-31058.

Morosinotto, T., M. Mozzo, et al. (2005). "Pigment-pigment interactions in Lhca4 antenna complex of higher plants photosystem I." Journal of Biological Chemistry 280(21): 20612-20619.

Mozzo, M., M. Mantelli, et al. (2010). "Functional analysis of Photosystem I light-harvesting complexes (Lhca) gene products of Chlamydomonas reinhardtii." Biochimica Et Biophysica Acta 1797(2): 212-221.

Muh, F., T. Renger, et al. (2008). "Crystal structure of cyanobacterial photosystem II at 3.0 A resolution: a closer look at the antenna system and the small membrane-intrinsic subunits." Plant Physiol Biochem 46(3): 238-264.

Mulkidjanian, A. Y., E. V. Koonin, et al. (2006). "The cyanobacterial genome core and the origin of photosynthesis." Proc Natl Acad Sci U S A 103(35): 13126-13131.

Muller, P., X. P. Li, et al. (2001). "Non-photochemical quenching. A response to excess light energy." Plant Physiology 125(4): 1558-1566.

Munekage, Y., M. Hashimoto, et al. (2004). "Cyclic electron flow around photosystem I is essential for photosynthesis." Nature 429(6991): 579-582.

Munekage, Y., M. Hojo, et al. (2002). "PGR5 is involved in cyclic electron flow around photosystem I and is essential for photoprotection in Arabidopsis." Cell 110(3): 361-371.

Murata, N. (1969). "Control of excitation transfer in photosynthesis. I. Light-induced change of chlorophyll a fluorescence in Porphyridium cruentum." Biochimica Et Biophysica Acta 172(2): 242-251.

Nelson, N. and A. Ben-Shem (2004). "The complex architecture of oxygenic photosynthesis." Nat Rev Mol Cell Biol 5(12): 971-982.

Nelson, N. and A. Ben-Shem (2005). "The structure of photosystem I and evolution of photosynthesis." Bioessays 27(9): 914-922.

Nelson, N. and C. F. Yocum (2006). "Structure and function of photosystems I and II." Annu Rev Plant Biol 57: 521-565.

Nield, J., M. Balsera, et al. (2002). "Three-dimensional electron cryo-microscopy study of the extrinsic domains of the oxygen-evolving complex of spinach: assignment of the PsbO protein." Journal of Biological Chemistry 277(17): 15006-15012.

Page 38: University of Groningen From Photosystem I to Photosystem

37

Nield, J., O. Kruse, et al. (2000). "Three-dimensional structure of Chlamydomonas reinhardtii and Synechococcus elongatus photosystem II complexes allows for comparison of their oxygen-evolving complex organization." Journal of Biological Chemistry 275(36): 27940-27946.

Nield, J., E. V. Orlova, et al. (2000). "3D map of the plant photosystem II supercomplex obtained by cryoelectron microscopy and single particle analysis." Nat Struct Biol 7(1): 44-47.

Nisbet, E. G. and N. H. Sleep (2001). "The habitat and nature of early life." Nature 409(6823): 1083-1091.

Niyogi, K. K. (1999). "PHOTOPROTECTION REVISITED: Genetic and Molecular Approaches." Annu Rev Plant Physiol Plant Mol Biol 50: 333-359.

Niyogi, K. K., O. Bjorkman, et al. (1997). "Chlamydomonas Xanthophyll Cycle Mutants Identified by Video Imaging of Chlorophyll Fluorescence Quenching." Plant Cell 9(8): 1369-1380.

Niyogi, K. K., O. Bjorkman, et al. (1997). "The roles of specific xanthophylls in photoprotection." Proc Natl Acad Sci U S A 94(25): 14162-14167.

Niyogi, K. K., A. R. Grossman, et al. (1998). "Arabidopsis mutants define a central role for the xanthophyll cycle in the regulation of photosynthetic energy conversion." Plant Cell 10(7): 1121-1134.

Niyogi, K. K., X. P. Li, et al. (2005). "Is PsbS the site of non-photochemical quenching in photosynthesis?" Journal of Experimental Botany 56(411): 375-382.

Nugent, J. H. (1996). "Oxygenic photosynthesis. Electron transfer in photosystem I and photosystem II." European Journal of Biochemistry 237(3): 519-531.

Okegawa, Y., T. A. Long, et al. (2007). "A balanced PGR5 level is required for chloroplast development and optimum operation of cyclic electron transport around photosystem I." Plant and Cell Physiology 48(10): 1462-1471.

Olson, J. M. (1981). "Evolution of photosynthetic and respiratory prokaryotes and organelles." Ann N Y Acad Sci 361: 8-19.

Olson, J. M. and R. E. Blankenship (2004). "Thinking about the evolution of photosynthesis." Photosynthesis Research 80(1-3): 373-386.

Pan, X. W., M. Li, et al. (2011). "Structural insights into energy regulation of light-harvesting complex CP29 from spinach." Nat Struct Mol Biol 18(3): 309-U394.

Peers, G., T. B. Truong, et al. (2009). "An ancient light-harvesting protein is critical for the regulation of algal photosynthesis." Nature 462(7272): 518-521.

Peltier, G., D. Tolleter, et al. (2010). "Auxiliary electron transport pathways in chloroplasts of microalgae." Photosynthesis Research 106(1-2): 19-31.

Peter, G. F. and J. P. Thornber (1991). "Biochemical-Evidence That the Higher-Plant Photosystem-Ii Core Complex Is Organized as a Dimer." Plant and Cell Physiology 32(8): 1237-1250.

Pribil, M., P. Pesaresi, et al. (2010). "Role of plastid protein phosphatase TAP38 in LHCII dephosphorylation and thylakoid electron flow." Plos Biology 8(1): e1000288.

Radmer, R. and B. Kok (1975). "Energy capture in photosynthesis: photosystem II." Annu Rev Biochem 44: 409-433.

Riesle, J., D. Oesterhelt, et al. (1996). "D38 is an essential part of the proton translocation pathway in bacteriorhodopsin." Biochemistry 35(21): 6635-6643.

Rochaix, J. D. (2001). "Assembly, function, and dynamics of the photosynthetic machinery in Chlamydomonas reinhardtii." Plant Physiology 127(4): 1394-1398.

Rochaix, J. D. (2002). "Chlamydomonas, a model system for studying the assembly and dynamics of photosynthetic complexes." Febs Letters 529(1): 34-38.

Roose, J. L., Y. Kashino, et al. (2007). "The PsbQ protein defines cyanobacterial Photosystem II complexes with highest activity and stability." Proc Natl Acad Sci U S A 104(7): 2548-2553.

Roose, J. L., K. M. Wegener, et al. (2007). "The extrinsic proteins of Photosystem II." Photosynthesis Research 92(3): 369-387.

Ruban, A. V., R. Berera, et al. (2007). "Identification of a mechanism of photoprotective energy dissipation in higher plants." Nature 450(7169): 575-578.

Page 39: University of Groningen From Photosystem I to Photosystem

38

Rupprecht, J. (2009). "From systems biology to fuel-Chlamydomonas reinhardtii as a model for a systems biology approach to improve biohydrogen production." Journal of Biotechnology 142(1): 10-20.

Scheibe, R., J. E. Backhausen, et al. (2005). "Strategies to maintain redox homeostasis during photosynthesis under changing conditions." Journal of Experimental Botany 56(416): 1481-1489.

Schopf, J. W. (1978). "The evolution of the earliest cells." Sci Am 239(3): 110-112, 114, 116-120 passim.

Schubert, W. D., O. Klukas, et al. (1998). "A common ancestor for oxygenic and anoxygenic photosynthetic systems: a comparison based on the structural model of photosystem I." J Mol Biol 280(2): 297-314.

Shapiguzov, A., B. Ingelsson, et al. (2010). "The PPH1 phosphatase is specifically involved in LHCII dephosphorylation and state transitions in Arabidopsis." Proc Natl Acad Sci U S A 107(10): 4782-4787.

Shi, L. X., M. Hall, et al. (2012). "Photosystem II, a growing complex: updates on newly discovered components and low molecular mass proteins." Biochimica Et Biophysica Acta 1817(1): 13-25.

Shi, L. X. and W. P. Schroder (2004). "The low molecular mass subunits of the photosynthetic supracomplex, photosystem II." Biochimica Et Biophysica Acta 1608(2-3): 75-96.

Shikanai, T. (2007). "Cyclic electron transport around photosystem I: genetic approaches." Annu Rev Plant Biol 58: 199-217.

Smith, B. M., P. J. Morrissey, et al. (1990). "Response of the Photosynthetic Apparatus in Dunaliella salina (Green Algae) to Irradiance Stress." Plant Physiology 93(4): 1433-1440.

Staehelin, L. A. (2003). "Chloroplast structure: from chlorophyll granules to supra-molecular architecture of thylakoid membranes." Photosynthesis Research 76(1-3): 185-196.

Standfuss, R., A. C. T. van Scheltinga, et al. (2005). "Mechanisms of photoprotection and nonphotochemical quenching in pea light-harvesting complex at 2.5A resolution." Embo Journal 24(5): 919-928.

Stauber, E. J., A. Busch, et al. (2009). "Proteotypic profiling of LHCI from Chlamydomonas reinhardtii provides new insights into structure and function of the complex." Proteomics 9(2): 398-408.

Swiatek, M., R. E. Regel, et al. (2003). "Effects of selective inactivation of individual genes for low-molecular-mass subunits on the assembly of photosystem II, as revealed by chloroplast transformation: the psbEFLJoperon in Nicotiana tabacum." Molecular Genetics and Genomics 268(6): 699-710.

Taiz, L. and E. Zeiger (2006). Plant Physiology, Sinauer Associates, Inc. Takahashi, H., M. Iwai, et al. (2006). "Identification of the mobile light-harvesting complex II

polypeptides for state transitions in Chlamydomonas reinhardtii." Proceedings of the National Academy of Sciences of the United States of America 103(2): 477-482.

Takahashi, Y., T. A. Yasui, et al. (2004). "Comparison of the subunit compositions of the PSI-LHCI supercomplex and the LHCI in the green alga Chlamydomonas reinhardtii." Biochemistry 43(24): 7816-7823.

Teramoto, H., T. Ono, et al. (2001). "Identification of Lhcb gene family encoding the light-harvesting chlorophyll-a/b proteins of photosystem II in Chlamydomonas reinhardtii." Plant and Cell Physiology 42(8): 849-856.

Thornton, L. E., H. Ohkawa, et al. (2004). "Homologs of plant PsbP and PsbQ proteins are necessary for regulation of photosystem ii activity in the cyanobacterium Synechocystis 6803." Plant Cell 16(8): 2164-2175.

Tokutsu, R., M. Iwai, et al. (2009). "CP29, a Monomeric Light-harvesting Complex II Protein, Is Essential for State Transitions in Chlamydomonas reinhardtii." Journal of Biological Chemistry 284(12): 7777-7782.

Page 40: University of Groningen From Photosystem I to Photosystem

39

Tokutsu, R., N. Kato, et al. (2012). "Revisiting the Supramolecular Organization of Photosystem II in Chlamydomonas reinhardtii." Journal of Biological Chemistry 287(37): 31574-31581.

Tokutsu, R. and J. Minagawa (2013). "Energy-dissipative supercomplex of photosystem II associated with LHCSR3 in Chlamydomonas reinhardtii." Proc Natl Acad Sci U S A 110(24): 10016-10021.

Umena, Y., K. Kawakami, et al. (2011). "Crystal structure of oxygen-evolving photosystem II at a resolution of 1.9 A." Nature 473(7345): 55-60.

Unlu, C., B. Drop, et al. (2014). "State transitions in Chlamydomonas reinhardtii strongly modulate the functional size of photosystem II but not of photosystem I." Proc Natl Acad Sci U S A.

Vallon, O., L. Bulte, et al. (1991). "Lateral Redistribution of Cytochrome B6/F Complexes Along Thylakoid Membranes Upon State Transitions." Proceedings of the National Academy of Sciences of the United States of America 88(18): 8262-8266.

van Grondelle, R., J. P. Dekker, et al. (1994). "Energy transfer and trapping in photosynthesis." Biochim. Biophys. Acta 1187(1-65).

van Roon, H., J. F. L. van Breemen, et al. (2000). "Solubilization of green plant thylakoid membranes with n-dodecyl-alpha,D-maltoside. Implications for the structural organization of the Photosystem II, Photosystem I, ATP synthase and cytochrome b(6)f complexes." Photosynthesis Research 64(2-3): 155-166.

Veeranjaneyulu, K. and R. M. Leblanc (1994). "Action Spectra of Photosystems I and II in State 1 and State 2 in Intact Sugar Maple Leaves." Plant Physiology 104(4): 1209-1214.

Vener, A. V., P. J. Van Kan, et al. (1995). "Activation/deactivation cycle of redox-controlled thylakoid protein phosphorylation. Role of plastoquinol bound to the reduced cytochrome bf complex." Journal of Biological Chemistry 270(42): 25225-25232.

Vener, A. V., P. J. M. VanKan, et al. (1997). "Plastoquinol at the quinol oxidation site of reduced cytochrome bf mediates signal transduction between light and protein phosphorylation: Thylakoid protein kinase deactivation by a single-turnover flash." Proceedings of the National Academy of Sciences of the United States of America 94(4): 1585-1590.

Wientjes, E. and R. Croce (2011). "The light-harvesting complexes of higher-plant Photosystem I: Lhca1/4 and Lhca2/3 form two red-emitting heterodimers." Biochemical Journal 433(3): 477-485.

Wientjes, E., G. T. Oostergetel, et al. (2009). "The role of Lhca complexes in the supramolecular organization of higher plant photosystem I." Journal of Biological Chemistry 284(12): 7803-7810.

Wientjes, E., H. van Amerongen, et al. (2013). "LHCII is an antenna of both photosystems after long-term acclimation." Biochimica Et Biophysica Acta 1827(3): 420-426.

Xiong, J. and C. E. Bauer (2002). "Complex evolution of photosynthesis." Annu Rev Plant Biol 53: 503-521.

Yachandra, V. K., K. Sauer, et al. (1996). "Manganese Cluster in Photosynthesis: Where Plants Oxidize Water to Dioxygen." Chem Rev 96(7): 2927-2950.

Yakushevska, A. E., P. E. Jensen, et al. (2001). "Supermolecular organization of photosystem II and its associated light-harvesting antenna in Arabidopsis thaliana." European Journal of Biochemistry 268(23): 6020-6028.

Yakushevska, A. E., W. Keegstra, et al. (2003). "The structure of photosystem II in Arabidopsis: Localization of the CP26 and CP29 antenna complexes." Biochemistry 42(3): 608-613.

Yamori, W., N. Sakata, et al. (2011). "Cyclic electron flow around photosystem I via chloroplast NAD(P)H dehydrogenase (NDH) complex performs a significant physiological role during photosynthesis and plant growth at low temperature in rice." Plant Journal 68(6): 966-976.

Yoon, H. S., J. D. Hackett, et al. (2004). "A molecular timeline for the origin of photosynthetic eukaryotes." Mol Biol Evol 21(5): 809-818.

Zhang, S. and H. V. Scheller (2004). "Light-harvesting complex II binds to several small subunits of photosystem I." Journal of Biological Chemistry 279(5): 3180-3187.

Page 41: University of Groningen From Photosystem I to Photosystem

40

Zito, F., G. Finazzi, et al. (1999). "The Qo site of cytochrome b(6)f complexes controls the activation of the LHCII kinase." Embo Journal 18(11): 2961-2969.

Zouni, A., H. T. Witt, et al. (2001). "Crystal structure of photosystem II from Synechococcus elongatus at 3.8 angstrom resolution." Nature 409(6821): 739-743.

Page 42: University of Groningen From Photosystem I to Photosystem

41

CHAPTER 2

Photosystem I of Chlamydomonas reinhardtii is composed of nine

Light-harvesting complexes (Lhca) located on one side of the core

Bartlomiej Drop1,4, Mariam Webber-Birungi1, Fabrizia Fusetti2, Roman Kouřil1, Kevin E.

Redding3, Egbert J. Boekema1 and Roberta Croce1,4*

1 Department of Biophysical Chemistry, Groningen Biological Sciences and Biotechnology Institute, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands;

2 Department of Biochemistry and Netherlands Proteomics Centre (NPC), Groningen Biological Sciences and Biotechnology Institute, University of Groningen, Nijenborgh 4 or 7, 9747 AG Groningen, The Netherlands;

3 Department of Chemistry and Biochemistry, Arizona State University, 1711 S. Rural Road, Tempe, AZ 85287-1604;

4 Department of Physics and Astronomy, Faculty of Sciences, VU University Amsterdam, De Boelelaan 1081, 1081 HV Amsterdam, The Netherlands.

Based on Journal of Biological Chemistry 2011, 286(52): 44878-44887

Page 43: University of Groningen From Photosystem I to Photosystem

42

Abstract

In this work we have purified the Photosystem I complex of Chlamydomonas

reinhardtii to homogeneity. Biochemical, proteomic, spectroscopic and structural analyses

reveal the main properties of this PSI-LHCI supercomplex. The data show that the largest

purified complex is composed of one core complex and nine Lhca antennas and that it

contains all Lhca gene products. A projection map at 15 Å resolution obtained by electron

microscopy reveals that the Lhcas are organized on one side of the core in a double half-ring

arrangement, in contrast with previous suggestions. A series of stable disassembled

intermediates of PSI-LHCI was purified. The analysis of these complexes suggests the

sequence of the assembly/disassembly process. It is shown that PSI-LHCI of C. reinhardtii is

larger, but far less stable, than the complex from higher plants. Lhca2 and Lhca9 (the red-

most antenna complexes), although present in the largest complex in 1:1 ratio with the

core, are only loosely associated with it. This can explain the large variation in antenna

composition of PSI-LHCI from C. reinhardtii found in the literature. The analysis of several

subcomplexes with reduced antenna size allows determination of the position of Lhca2 and

Lhca9 and leads to a proposal for a model of the organization of the Lhcas within the PSI-

LHCI supercomplex.

Page 44: University of Groningen From Photosystem I to Photosystem

43

1. Introduction

The light reactions of photosynthesis are carried out by four large protein complexes

– Photosystem II (PSII), cytochrome b6f, Photosystem I (PSI) and ATP-synthase. Like the

other complexes, PSI is located in the thylakoid membrane of cyanobacteria, algae and

plants (Arnon 1971; Dekker and Boekema 2005). It catalyzes light-driven electron transport

from plastocyanin, which is present in the thylakoid lumen, to ferredoxin, which is located in

the stroma. With almost 100% quantum efficiency, PSI is considered to be the most efficient

light capturing and energy conversion apparatus found in Nature (Jensen, Haldrup et al.

2003; Amunts and Nelson 2009).

PSI evolved well over 1 billion years ago, probably over 2 billion years ago. It is thought to

represent an elaboration of the simple homodimeric type I reaction centers of primordial

organisms into the sophisticated machinery that we know from plants and algae (Amunts

and Nelson 2009). During evolution, different groups of photosynthetic organisms colonized

diverse ecological niches. As a consequence of adaptation to different environmental

conditions, PSI design underwent structural rearrangements, most having to do with the

peripheral antenna system. In the lineage that produced green algae and plants, this led to

the addition of several specific complexes that enhance the light-harvesting capacity of the

system (Ben-Shem, Frolow et al. 2004; Nelson and Ben-Shem 2005; Amunts and Nelson

2008). In cyanobacteria PSI exists as a trimeric (Boekema, Dekker et al. 1987), while in plants

and green algae, PSI is monomeric and is composed of two moieties: the PSI core complex

and an outer antenna system composed of light harvesting complexes (LHCI) (Lam, Oritz et

al. 1984).

The PSI core is responsible for light-driven charge separation and electron transfer.

It coordinates around 100 chlorophylls and 20 β-carotene molecules. Its primary and tertiary

structures are highly conserved among green algae and plants: 14 subunits are present in

both types of organisms (PsaA – PsaL and PsaN – PsaO), while PsaP is present in plants but

so far seems to be absent in algae (Jensen, Haldrup et al. 2003; Ozawa, Onishi et al. 2010).

The central part of the core complex is composed of two large proteins – PsaA and PsaB,

which binds all of the cofactors of the electron transfer chain (Jordan, Fromme et al. 2001),

except for the last 2 Fe4S4 clusters (FA and FB). These are bound to the peripheral subunit

PsaC, which together with PsaD and PsaE forms the docking site for ferredoxin, on the

stromal side of the membrane (Scheller, Jensen et al. 2001; Jensen, Bassi et al. 2007). PsaF

Page 45: University of Groningen From Photosystem I to Photosystem

44

and PsaN are important for electron transfer from plastocyanin to P700 (Farah, Rappaport

et al. 1995; Haldrup, Naver et al. 1999). PsaJ is a hydrophobic protein located close to PsaF

and plays a role in the stabilization of this subunit’s conformation (Fischer, Boudreau et al.

1999). PsaH, PsaI, PsaL, PsaO form a cluster of integral membrane proteins, placed on one

side of the core, where they are involved in interactions with LHCII during state transitions

(Lunde, Jensen et al. 2000; Zhang and Scheller 2004). PsaG and PsaK are located near PsaA

and PsaB respectively (Scheller, Jensen et al. 2001; Jensen, Bassi et al. 2007) and have been

proposed to be important for the association of the outer antenna with the core (Jensen,

Gilpin et al. 2000; Ben-Shem, Frolow et al. 2003; Ozawa, Onishi et al. 2010). The outer

antenna extends the light-harvesting capacity and ensures photoprotection (Ben-Shem,

Frolow et al. 2003). The genes encoding the antenna complexes of PSI of plants and green

algae are members of the Lhc multigene family, which also includes the antenna complexes

of PSII. These complexes show structural homology in that each Lhc polypeptide has three

transmembrane α-helices and coordinates Chls a, Chls b, and different carotenoid molecules

(Jansson 1999).

The structure of PSI-LHCI in land plants has been determined to 3.3 Å resolution and reveals

the location of four Lhca proteins (encoded by the lhca 1-4 genes) on one side of the core

(Scheller, Jensen et al. 2001; Ben-Shem, Frolow et al. 2003; Amunts, Drory et al. 2007). Two

additional Lhca genes, lhca5 and lhca6, were identified in Arabidopsis thaliana (A.t.), but

their products are present at much lower levels compared with the other Lhca complexes

(Ganeteg, Klimmek et al. 2004; Storf, Jansson et al. 2005; Lucinski, Schmid et al. 2006). The

structure of the PSI-LHCI complex of green algae is not available, but biochemical and

structural (via electron-microscopy) analyses have suggested that PSI-LHCI of

Chlamydomonas reinhardtii differs in size and antenna composition from its counterpart in

plants (Germano, Yakushevska et al. 2002; Kargul, Nield et al. 2003). Indeed, 9 lhca genes

have been identified in C. reinhardtii (Elrad and Grossman 2004). All of them are expressed

in normal conditions (Stauber, Fink et al. 2003), and all gene products were shown to

coordinate pigments (Mozzo, Mantelli et al. 2010). Based on their content of long-

wavelength Chls (‘red forms’) and on their fluorescence emission maxima, the Lhca

complexes were divided into three subclasses: “blue Lhca” (Lhca1, Lhca3 and Lhca7) with

emission maxima at 682.5–683.5 nm, “intermediate Lhca” (Lhca5, Lhca6 and Lhca8) with

Page 46: University of Groningen From Photosystem I to Photosystem

45

maxima between 694.5 and 697.5 and “red Lhca” (Lhca2, Lhca4 and Lhca9) with maxima

between 707 and 715 nm (Mozzo, Mantelli et al. 2010).

Based upon biochemical characterization, it was suggested that C. reinhardtii PSI-LHCI

(C.r.PSI-LHCI) contains between 6 and 14 Lhca subunits (Bassi, Soen et al. 1992; Takahashi,

Yasui et al. 2004; Tokutsu, Teramoto et al. 2004; Subramanyam, Jolley et al. 2006). Recently

the Lhca/core stoichiometry was estimated using mass spectrometry, suggesting the

presence of 7.5 ± 1.4 Lhca subunits per PSI-core (Stauber, Busch et al. 2009). The Lhca

complexes were categorized into three groups: those present in a substoichiometric amount

with respect to the core (Lhca2, Lhca5, Lhca6, Lhca8 and Lhca9), those present in a 1:1 ratio

with the core (Lhca1, Lhca4 and Lhca7), and Lhca3 present at a 2:1 ratio (Stauber, Busch et

al. 2009), suggesting a heterogeneous composition of the PSI-LHCI complex. It was also

shown that six Lhca subunits (Lhca1, Lhca4–8) may even assemble in the absence of PSI core

complex (Wollman and Bennoun 1982; Takahashi, Yasui et al. 2004).

Three structural models of C.r.PSI-LHCI complex have been proposed based on single

particle electron microscopy analysis (Germano, Yakushevska et al. 2002; Kargul, Nield et al.

2003; Kargul, Turkina et al. 2005). When compared with higher plant PSI, the C.r.PSI-LHCI

particles displayed additional densities, suggesting a larger antenna size. Germano et al.

(2002) suggested that about 14 Lhca complexes are associated with the core. Most of them

are present on the PsaJ/PsaF side, while 2-3 Lhca bind to the PsaH subunit. Kargul and

coworkers, proposed that 11 Lhcas are arranged on the PsaG/F/J/K side (Kargul, Nield et al.

2003), but they later revised their model, suggesting that LHCI is made up of 6 Lhca

subunits, four of which are located at the same position as the four Lhcas in higher plants,

and two of which are at the PsaH side (Kargul, Turkina et al. 2005).

In this work we have purified several PSI (sub)complexes that differ in protein composition

and we have analyzed them via a combination of biochemical, spectroscopic and electron-

microscopy (EM) measurements, with the aim of revealing the functional architecture of

this complex.

2. Materials and Methods

2.1. Strain and growth conditions

A C. reinhardtii strain (JVD-1B[pGG1]) in which a hexa-histidine tag has been added to

the N-terminus of PsaA (Gulis, Narasimhulu et al. 2008) was grown in liquid Tris-Acetate-

Page 47: University of Groningen From Photosystem I to Photosystem

46

Phosphate (TAP) medium (Gorman and Levine 1965). Cells were cultured at room

temperature (25°C) on a rotary shaker under an illumination at 10 μmol photons PAR m-2 s-1.

2.2. Thylakoids preparations

Cells were harvested by centrifugation (4000 rpm, 6 min, 4°C) in the mid-logarithmic phase

(OD750nm ≈ 0.7) and thylakoids were prepared as in Fischer et al. (1997) with a few

modifications. Cells were disrupted by sonication (60W power in 10 cycles of 10 s on/30 s

off) and centrifuged (15000 rpm, 20 min, 4°C). Thylakoid membranes were separated on

a discontinuous gradient (24000 rpm, 1 h, 4°C) in a SW41 swinging bucket rotor as described

(1997).

2.3 Purification of the His-tagged PSI complex

Thylakoid membranes were washed three times with a buffer containing 10 mM Hepes

(pH 7.5) and 5 mM EDTA. Membranes were resuspended in solubilization buffer (20 mM

Hepes (pH 7.5), 0.2 M NaCl) to a chlorophyll concentration of 1 mg/ml and an equal volume

of 0.9% α-dodecylmaltoside (α-DM) was added. Samples were vortex for few seconds and

then centrifuged (15000 rpm, 10 min, 4 °C) to remove unsolubilized material.

The supernatant was loaded onto a HisTrap HP Column (GE Healthcare). The column was

equilibrated with 20 mM Hepes (pH 7.5), 0.2 M NaCl, 0.03% α-DM and washed with the

same buffer. Elution was carried out with 20 mM Hepes (pH 7.5), 0.25 M imidazole, 0.03%

α-DM. The eluted fraction was loaded on a sucrose density gradient (made by freezing and

thawing 0.65 M sucrose, 10 mM Tricine (pH 7.8), 0.03% α-DM buffer) and purified to

homogeneity by ultracentrifugation (41000 rpm, 17 h, 4 ºC).

The purification of PSI from A.t., was performed under the same conditions used for

the purification of the Chlamydomonas PSI, with the difference that upon solubilization

the thylakoid membranes were directly loaded on the sucrose gradient.

2.4 Gel electrophoresis

Proteins were analyzed by a SDS-PAGE in a Tris-sulphate buffer system prepared as in Bassi

(1985). 3 µg of sample (in Chls) was loaded into each well. Non-denaturing gels were

prepared as in Peter and Thornber (1991) with a few modifications. The resolving gel

contained a 4-10% acrylamide gradient and 0.01% SDS was added to the upper running

buffer. Some of the PSI-LHCI samples were solubilised with 1% β-DM and 0.4%

ZWITTERGENT (at a Chl concentration of 0.4 mg ml-1).

Page 48: University of Groningen From Photosystem I to Photosystem

47

2.5 In-gel tryptic digestion

For mass spectrometry-based protein identification, the green bands were excised from the

gel and treated with 10 mM DTT followed by 55 mM iodoacetamide in 50 mM NH4HCO3, to

reduce and alkylate cysteine residues, and subsequently dehydrated by incubation for 5 min

in 100% acetonitrile. The gel slices were rehydrated in 10 µl trypsin solution (Trypsin Gold,

Mass spectrometry grade, Promega 10 ng/µl in 25 mM NH4HCO3), and incubated for 2 hours

at 37°C. Subsequently 10 µl of 25 mM NH4HCO3 was added to prevent drying and

the incubation was prolonged overnight at 37°C. The tryptic peptides were recovered by

three subsequent extractions with 50 µl of 35%, 50% and 70% acetonitrile in 0.1% TFA.

The extracted peptides were pooled and concentrated under vacuum.

For in-solution tryptic digestion, samples containing around 10 µg of protein were diluted to

20 µl in 100 mM NH4HCO3. For reduction and alkylation, the samples were incubated for 60

min at 55°C in the presence of 10 mM DTT, followed by addition of iodoacetamide (final

concentration of 50 mM) and incubation at room temperature for 30 min. After addition of

0.5 µg of trypsin, the samples were incubated over night at 37°C.

2.6 Liquid Chromatography-Mass Spectrometry (LC-MS)

Fractions of the peptide mixtures from in-gel trypsin digestions were diluted in 5% formic

acid, passed through a pre-column (EASY-Column C18, 100 µm x 20mm, 5µm particle size,

Thermo Scientific, Bremen, Germany) and separated on a capillary column (C18 PepMap

300, 75 µm x 250 mm, 3-µm particle size, Dionex, Amsterdam, The Netherlands) mounted

on a Proxeon Easy-LC system (Proxeon Biosystems, Odense, Denmark). Solutions of 0.1%

formic acid in water and a 0.1% formic acid in 100% acetonitrile were used as the mobile

phases. A gradient from 4 to 40% acetonitrile was performed in 140 min, at a flow rate of

300 nL/min. Eluted peptides were analyzed using a linear ion trap-orbitrap hybrid mass

spectrometer (LTQ-Orbitrap, Thermo Scientific). MS scans were acquired in the Orbitrap, in

the range from 400 to 1800 m/z, with a resolution of 60,000 (FWHM). The 7 most intense

ions per scan were submitted to MS/MS fragmentation (35% Normalized Collision EnergyTM)

and detected in the linear ion trap.

2.7 Protein identification

The MS raw data were submitted to Mascot (version 2.1, Matrix Science, London, UK) using

the Proteome Discoverer 1.0 analysis platform (Thermo Scientific) and searched against the

Page 49: University of Groningen From Photosystem I to Photosystem

48

C. reinhardtii proteome. Peptide tolerance was set to 10 ppm and 0.9 Da for intact peptides

and fragment ions respectively, using semi-trypsin as protease specificity and allowing for

up to 2 missed cleavages. Oxidation of methionine residues, deamidation of asparagine and

glutamine, and carboamidomethylation of cysteines were specified as variable

modifications. The MS/MS based peptide and protein identifications were further validated

with the program Scaffold (version Scaffold_3.0, Proteome Software Inc., Portland, OR).

Protein identifications based on at least 2 unique peptides identified by MS/MS, each with a

confidence of identification probability higher than 95%, were accepted. All assigned spectra

used for the spectral count analysis were derived from at least two experiments, with the

exception of the Lhca content in fraction A4.

2.8 Spectroscopic analysis

Room-temperature and 77K absorption spectra were recorded with a Cary 4000

spectrophotometer (Varian). Room-temperature and 77K fluorescence emission spectra

were recorded using a Fluorolog 3.22 spectrofluorimeter (Jobin Yvon-Spex). The excitation

wavelengths were 440 nm, 475 nm and 500 nm and excitation and emission slits were set to

3.5 nm. For low-temperature measurements, samples were in 66.7% glycerol (w/v), 10 mM

Tricine (pH 7.8), 0.03 % α-DM.

2.9 Pigment composition

The pigment composition of the complexes was analyzed by fitting the spectrum of the

acetone extracted pigments with the spectra of the individual pigments in acetone and by

HPLC as described (Croce, Canino et al. 2002). The data are the results of at least two

different preparations.

2.10 Electron Microscopy and Single Particle Analysis

To improve sample purity and contrast in electron-microscopy micrographs, the sucrose

content of the gradient fractions was reduced by dialysis (2-4 h at 4°C; Spectra/Por 12–14

kDa cut-off). After dialysis, specimens were prepared for negative staining with 2% uranyl

acetate on glow-discharged carbon-coated copper grids. Electron microscopy was

performed on a Philips CM120 electron microscope equipped with a LaB6 tip, operated at

120 kV. Images were recorded with a Gatan 4000 SP 4K slow-scan CCD camera at 133,000

magnification with a pixel size of 0.225 nm at the specimen level after binning the images.

Page 50: University of Groningen From Photosystem I to Photosystem

49

GRACE software was used for semi-automated data acquisition (Oostergetel, Keegstra et al.

1998). Single particles were analyzed with the Groningen Image Processing (GRIP) software,

including multi-reference and non-reference alignments, multivariate statistical analysis and

classification, as previously described (1999). To sharpen the final images, limited numbers

of projections (500-2000) of homogeneous classes were summed, with the correlation

coefficient in the final alignment step as a quality parameter.

3. Results

3.1 Isolation and fractionation of PSI-LHCI complexes

PSI-LHCI was purified in two steps from a strain carrying a hexa-histidine tag at

the N-terminus of core subunit PsaA. Thylakoid membranes were mildly solubilized and PSI

particles were first purified on a Ni-Sepharose column. To obtain homogeneous

preparations of PSI and to check for the presence of PSI complexes with different sizes, the

eluted fraction was then subjected to sucrose density gradient ultracentrifugation. Two

green bands were separated (B1 and B2; Fig. 1A), indicating that the elution contained two

populations of complexes differing in molecular weight. These chlorophyll-protein

complexes were collected and subjected to biochemical and spectroscopic analysis.

Figure 1. A: Fractionation of PSI-LHCI eluted from the Ni(II)-column by sucrose density gradient. B: SDS-PAGE of green bands B1 and B2 from the gradient. Panel C: Venn diagram showing the composition of the PSI-LHCI preparation (fraction B2) based on LC-MS/MS analysis of tryptic peptides. The sum of the spectra assigned to each group of proteins, as well as the fraction of total spectra, is indicated.

3.2 PSI-LHCI: Protein composition

The protein compositions of the two fractions were visualized by SDS-PAGE (Fig. 1B).

The polypeptide profile analysis confirmed the presence of PSI in both fractions. A broad

Page 51: University of Groningen From Photosystem I to Photosystem

50

band at high molecular weight corresponds to the PSI reaction centre core subunits, PsaA

and PsaB, which have an apparent molecular weight of ~65 kDa, and typically migrate as

a diffuse band. Bands in the 20-27 kDa range correspond to LHCI antennae and the stromal

subunits PsaD and PsaF, while those in the 10-20 kDa range arise from small core subunits

(Bassi, Soen et al. 1992; Ozawa, Onishi et al. 2010). No bands corresponding to PSII subunits

were visible in the gel, indicating that both PSI complexes were free from PSII

contamination, as seen before with the unfractionated eluate (Gulis, Narasimhulu et al.

2008). While the band pattern was very similar for the two fractions, the ratio between core

subunits and Lhca complexes was higher in B1 than in B2, suggesting that the two fractions

contain PSI complexes with different antenna sizes. B2 thus represents the largest complex

and in the following it is called PSI-LHCI. B1 is likely the result of a partial disassembly of PSI-

LHCI during purification, as indicated by the fact that Lhca complexes were observed in the

flow-through of the Ni-column (Fig. 2). However, we cannot exclude that this fraction also

contains the assembly intermediate recently observed by Ozawa et al. (2010).

Figure 2. SDS-PAGE analysis of the gradient bands obtained from the “flow-though” fraction of the Ni(II)-column. The presence of two PSI supercomplexes with different size in the sucrose gradient following affinity chromatography suggested that part of the antenna has been lost during purification. To test this hypothesis, the fraction that did not bind to the column (“flow-through’) was loaded on a sucrose density gradient, and the green bands (not shown) were harvested and analysed by SDS-PAGE. Lhca polypeptides were visible in the upper bands, along with antenna complexes of PSII (Lhcb), indicating that indeed some of the Lhca complexes are lost during purification.

To determine in more detail the protein composition of PSI-LHCI, we performed a mass

spectrometry analysis. A shotgun proteomics approach revealed the presence of all nine

Lhca proteins (Lhca1–9) and all PSI core subunits, with the exception of PsaI and PsaO.

These subunits are likely to be present in the complex but unidentifiable by tryptic

Page 52: University of Groningen From Photosystem I to Photosystem

51

digestion, because the resulting peptides would be too short and would not be detectable in

our experimental setup. PsaN and PsaJ were not identified in all experiments, and were only

seen via one unique peptide, so they are not considered in the analysis.

The proteomic analysis also confirms the high purity of the preparation (Fig. 1C): 94% of the

MS/MS spectra were assigned to peptides originating from PSI or LHCI, with 55% from Lhca

polypeptides and 39% from PSI core subunits. Only 6% were due to impurities, of which 4%

stemmed from ATP synthase subunits. No PSII or LHCII subunits were detected in the

preparation.

3.3 PSI-LHCI: Pigment composition

Pigment analysis indicated a Chl a/b ratio of 4.4 in the purified PSI-LHCI complex (Table 1).

The main carotenoids were β-carotene, lutein and violaxanthin. Loroxanthin was present in

low amounts, in agreement with previous results indicating that this xanthophyll is mainly

associated with PSII (Pineau, Gerard-Hirne et al. 2001). No traces of neoxanthin could be

detected, consistent with the fact that this xanthophyll is not present in PSI and that the

preparation is not contaminated by PSII complexes.

Table 1. Pigment compositions of PSI-LHCI supercomplexes from C. reinhardtii (this study)

and A. thaliana (Wientjes, Oostergetel et al. 2009). The values of individual carotenoids are

normalized to 100 Chls (a+b).

Chl a/b Chl/car Loroxanthin Violaxanthin Lutein β-carotene

C.r.PSI-LHCI 4.4 ± 0.1 5.0 ± 0.2 2.0 ± 0.6 4.8 ± 0.6 7.0 ± 1.0 8.5 ± 0.3

A.t.PSI-LHCI 9.7 ± 0.3 4.8 ± 0.1 - 2.3 ± 0.3 5.5 ± 0.3 13.1 ± 0.3

The Chl a/b ratio of C.r.PSI-LHCI is far lower than that of the PSI-LHCI complex of higher

plants (a/b 8.5-9) (van Oort, Amunts et al. 2008; Wientjes, Oostergetel et al. 2009).

We envision two possible origins for this difference: (1) a higher number of antenna

complexes associated with C.r.PSI-LHCI as compared to plant PSI-LHCI, and/or (2) a higher

complement of Chl b in the Lhca complexes of C. reinhardtii as compared with the

homologues of higher plants. Both effects are probably contributing. Indeed, if the

chlorophyll composition of the C.r.Lhca were the same as in higher plants (14 Chls per

complex with a Chl a/b ratio of 3.7 (Wientjes and Croce 2011)), then 48 Lhcas would be

needed to explain a Chl a/b ratio of 4.4 in C.r.PSI-LHCI! On the other hand, to explain this

value purely in terms of a different affinity for Chl a and b, the C.r.Lhcas would need to have

Page 53: University of Groningen From Photosystem I to Photosystem

52

a Chl a/b ratio of 1.2, which is far lower than observed for both native and reconstituted

complexes (Bassi, Soen et al. 1992; Kargul, Nield et al. 2003; Takahashi, Yasui et al. 2004;

Mozzo, Mantelli et al. 2010).

3.4 PSI-LHCI: Spectroscopic characterization

The PSI-LHCI absorption spectrum at room temperature shows a maximum at 679 nm (Fig.

3A). Compared with the complex of higher plants (Fig. 3A), C.r.PSI-LHCI is characterized by (i)

an increased absorption in the Chl b region (630-660 nm), in agreement with the lower Chl

a/b ratio; and (ii) a decreased absorption in the region above 700 nm, indicating a lower

content of red forms, as observed previously (Bassi, Soen et al. 1992; Germano,

Yakushevska et al. 2002).

Figure 3. Room-temperature absorption (A) and 77 K fluorescence emission (B and C) spectra ofC.r.PSI-LHCI (A and C,solid line; B, dashed line) compared with those of cells (panel B, dotted line) andA.t.PSI-LHCI (panels A and C,dashed line). Excitation used for fluorescence spectra was at 440 nm. The difference between normalized fluorescence emission spectra of cells and C.r.PSI-LHCI is shown inpanel B (solid line).

The fluorescence emission spectrum of C.r.PSI-LHCI at low temperature (77K) exhibits

a maximum at 715 nm and it is practically identical to the spectrum of PSI in the cell

(Fig. 3B), indicating that the purified complex maintains its in vivo properties. Although

the emission maximum of the C. reinhardtii complex is 18 nm blue-shifted compared to

the spectrum of plant PSI-LHCI (Fig. 3C and (Kargul, Nield et al. 2003; Takahashi, Yasui et al.

Page 54: University of Groningen From Photosystem I to Photosystem

53

2004; Gibasiewicz, Szrajner et al. 2005; Gulis, Narasimhulu et al. 2008)), the two spectra

have similar characteristics. They are very broad and strongly red-shifted, with respect to

the absorption of the complexes, suggesting a similar origin of the red forms (i.e. mixing of

the lowest excitonic state with a charge transfer state (Ihalainen, Ratsep et al. 2003;

Romero, Mozzo et al. 2009)), in the two organisms.

3.5 Native PAGE of PSI-LHCI subcomplexes

In order to investigate the stability of PSI-LHCI and to assess the strength of the association

of the individual Lhca complexes with the PSI core, PSI-LHCI was treated with a combination

of Zwittergent and β-DM, and the resulting subcomplexes were separated by native PAGE.

Five bands (A1-A5; Fig. 4A) were observed. Surprisingly, even a very low concentration of

ionic detergent led to a partial dissociation of the complex, as indicated by the separation of

PSI-LHCI into two main bands in 0.01% SDS (Fig. 4B), suggesting that part of the antenna is

associated with the core in a relatively loose fashion.

Figure 4. Native PAGE of PSI-LHCI supercomplexes of C. reinhardtii (Cr) or A. thaliana (At) solubilized with 1% β-DM and 0.4% Zwittergent-16 (A) or with 0.01% SDS (B). C: Native gel of the B1 band solubilized with 0.01% SDS (Cr) compared with the PSI-LHCI (sub)complexes of Arabidopsis thaliana (At).

The molecular weight of the PSI (sub)complexes can be roughly estimated by using A.t.PSI-

LHCI (600 KDa) and A.t.PSI-core (440 KDa) particles as standards. Approximate values of 770,

690, 570, 460 and 370 kDa were found for the MWs of bands A1-A5, respectively. Assuming

an average MW of 35 kDa for an Lhca complex, the A1 fraction should therefore contain

a PSI complex with 5 Lhcas more than A.t.PSI-LHCI, and thus 9 in total. The A2 fraction

would therefore have lost 2 Lhcas compared to A1, while only 3-4 Lhcas are expected to be

present in fraction A3. A4 contains a complex smaller than A.t.PSI-LHCI but larger than

A.t.PSI-core, suggesting that it probably still binds 1-2 Lhcas. Fraction A5 migrates below the

Page 55: University of Groningen From Photosystem I to Photosystem

54

A.t.PSI-core, indicating that it had lost some of the core subunits. An additional complex,

fraction A6, with an apparent MW of 140 kDa, was obtained upon very mild solubilization of

the B1 band.

3.6 PSI (sub)complexes: Protein composition

To get information about the protein composition of the PSI complexes, the A1-A6 fractions

were analyzed by mass spectrometry. To determine the relative abundance of each subunit

in the different fractions, spectral count normalization was applied (Liu, Sadygov et al.

2004).

Figure 5. Quantification of the LHCI (A) and PSI core (B) polypeptides found in the PSI-LHCI subcomplexes. The number of spectra assigned to individual subunits was normalized to the number of spectra assigned to the PsaA/PsaB polypeptides in the same fraction. The values are reported relative to the values determined for each polypeptide in fraction A1.

The Lhca content from all bands is presented in Fig. 5A as relative values normalized to the

abundance of the same subunit in the largest complex. The A2 fraction shows a 90%

decrease of Lhca2 and Lhca9, while A3 shows an additional loss of 95% of Lhca4, Lhca5 and

Lhca6. Surprisingly, in fractions A4 and A5, for which the mobility on the gel suggests that

they represent PSI with strongly reduced antenna size, all Lhcas were detected, with the

exception of Lhca2 and Lhca9. However in fraction A4, while the amount of Lhca7 and Lhca8

is almost unchanged compared to A1, it is around 40% for all other Lhcas. In the case of

Page 56: University of Groningen From Photosystem I to Photosystem

55

fraction A5, all the Lhca polypeptides are present in similar amounts, but around 20-30% of

their content in A1. Virtually no Lhca proteins were detected in fraction A6.

The composition of the PSI core subunits in the different fractions was also analyzed

(Fig. 5B). The data indicate a strong reduction of PsaL and PsaH in fractions A2 and A3.

Interestingly, in fractions A4 and A5, PsaF together with PsaG and PsaK, which have been

suggested to stabilize the binding of Lhca complexes to the core, are significantly reduced.

In fraction A6 the stromal subunits PsaC, D and E are also missing; it still contains PsaA and

PsaB, and likely represents a core heterodimer lacking all small subunits or peripheral

antennae.

3.7 PSI (sub)complexes: Absorption, fluorescence and pigment analysis

The absorption maxima of the complexes shift to the blue as one descends to less and less

assembled states. While the absorption maximum of A1 is 679 nm, it is 677 nm in A4 and A5

(Fig. 6A). A decrease in the Chl b absorption region is also observed going from A1 to A5,

with the exception of A4, which shows higher absorption in this region than A3.

Figure 6. Absorption (A) and low-temperature fluorescence emission (B) spectra of the PSI-LHCI subcomplexes shown in Figure 4.

Fluorescence emission spectra were measured at 77K (Fig. 6B). The emission maximum is

715 nm for A1 and A2, 711 nm for A3, and 708 nm for A4 and A5 (Table 2). The second peak

around 680 nm observed in A3 is due to the presence of a pool of disconnected Chls in this

fraction.

Page 57: University of Groningen From Photosystem I to Photosystem

56

The pigment compositions of the five fractions were also analyzed and are reported in Table

2. An increase in Chl a/b ratio was observed in going from fraction A1 to A5, again with the

exception of A4, which shows a lower Chl a/b ratio as compared to band A3.

Table 2. Pigment composition of PSI-LHCI subcomplexes A1-A5. (The error in these

measurements is below 0.2.)

Chl a/b Chl/car Abs max (nm) Emission max (nm)

A1 4.4 ± 0.1 5.5 ± 0.4 679 715

A2 4.5 ± 0.1 5.6 ± 0.4 679 715

A3 5.7 ± 0.3 6.3 ± 1.7 678 711

A4 4.6 ± 0.2 6.7 ± 0.3 677 708

A5 6.5 ± 0.5 7.6 ± 1.6 677 708

3.8 Electron microscopy and single-particle analysis

To determine the structural organization of PSI-LHCI, electron microscopy and single-

particle analysis were performed. Specimens of PSI-LHCI yielded a relatively homogeneous

preparation (Fig. 7).

Figure 7. Electron micrograph of negatively-stained C.r.PSI-LHCI, demonstrating the homogeneity of the preparation. Scale bar = 100 nm.

Single-particle analysis of more than 50,000 projections gave two major classes, which

represent two types of complexes of a slightly different overall size. The larger PSI-LHCI

complex (Fig. 8A) represents about 40% of the dataset, while the smaller type (Fig. 8B)

accounts for about 55%. The main difference between the two particles is on the right side

of the projection map, with a clear absence of density in the smaller complex. The

difference is not a matter of a variation in tilt on the carbon support film (Fig. 9), but is likely

due to the loss of a specific subunit.

Page 58: University of Groningen From Photosystem I to Photosystem

57

Figure 8. Average projection maps of C.r.PSI-LHCI (A,B) and the A2 fraction (C). The maps in panels A and B depict the two major class averages of PSI-LHCI complex, representing sums of about 2000 particles each. The larger particle (A) represents ~40% of the population in the fraction, while the smaller one (B) represents ~55% of the particles. Panel C shows the sum of 500 particles of the Lhca2/Lhca9-less complex from fraction A2. The black and white arrows indicate the presence and absence of two densities of similar size and shape at the top. Space bar = 10 nm.

For a structural assignment of the PSI-LHCI projection maps, determination of their

handedness is crucial. This was facilitated by the presence of a characteristic crescent-

shaped density at the right side of the EM projection maps (Fig. 8), which is typical for the

PSI-LHCI complex seen from the stromal side (Fig. 10A and (Wientjes, Oostergetel et al.

2009)).

Figure 9. Analysis of the EM projection of PSI-LHCI and comparison of the two projections. Comparison between the projection maps of two main classes of particles observed in the PSI-LHCI fraction. A and B are the same as in Fig. 8. The contour of the particles and some internal high density areas have been marked in green and red. The overlay of the two profiles (C) shows that the positions of the high-density areas are very similar, indicating that the differences between the two particles are not caused by a different tilting of the same particle on the carbon support film used for EM specimen preparation. Hence, the difference in density is caused by the absence of a protein subunit in B compared to A. The extra area observed in A has a width of around 15 Å, and it is thus too small to accommodate an Lhca complex. It is suggested that the extra density represents PsaH and/or PsaL subunits in agreement with the biochemical data that shows that these subunits are easily lost during purification. Scale bar is 10 nm.

Page 59: University of Groningen From Photosystem I to Photosystem

58

Figure 10. Analysis of the EM projection of PSI-LHCI and comparison with plant supercomplexes. The EM projection maps of C.r.PSI-LHCI (A,B) and A.t.PSI-LHCI-LHCII supercomplexes (C, adapted from Kouřil et al. 2005) and a projection map of pea PSI-LHCI complex (D) viewed from the stromal side and generated from the atomic model (Amunts et al. 2010) by truncating to 15-Å resolution using routines from the EMAN package (Ludtke et al. 1999). The grey contours mark the strongest density of the PSI core complex. The white arrow in D indicates the strong density of the C helix of Lhca2 subunit, which was used as a marker for the assignment of the high-density spots of the outer Lhca half-ring in the C.r.PSI-LHCI supercomplex (see Figure 11). The C helix is almost perpendicular to the membrane plane and separated from the other two helices. Under such circumstances staining molecules can optimally accumulate around this helix and a projection of such a helix will result in a clear, strong density spot. Yellow arrows in (A) indicate high densities resolved in the outer row of Lhca, which is attributed to the C helix of Lhca protein in (the one of Lhca2 is indicated in D).

This characteristic density facilitated a manual fitting of the X-ray model of plant PSI-LHCI

(Amunts, Toporik et al. 2010) into the projection maps of the C.r.PSI-LHCI (Fig. 11A,B). The

fit shows an overall good agreement between X-ray and EM data, especially for the larger

complex (Fig. 11B). In the case of the smaller complex, the lack of EM density on the right

side of the projection map may correspond to the absence of PsaL and PsaH (Fig. 11, white

arrows), rather than to the absence or presence of a larger-sized Lhca subunit. This

assignment is supported by the biochemical data indicating that PsaH and PsaL are the first

subunits to be lost (see above). An on-scale comparison of the C.r.PSI-LHCI projection maps

with the structure of pea PSI-LHCI clearly indicates that the antenna proteins of C.reinhardtii

are all bound at one side of the core (Fig. 11A,B). To delve further into the organization of

the Lhca complexes, we performed a similar analysis of fraction A2, which lacks Lhca2 and

Lhca9. Electron microscopy of the A2 specimen, followed by image analysis of about 20,000

single-particle projections, produced a projection map similar to, but smaller than, that of

the C.r.PSI-LHCI complex (Fig. 8C). A direct comparison of the projection maps of PSI-LHCI

and A2 (Fig. 8 A and C) leads to the localization of Lhca2 and Lhca9. One of them occupies

the binding site where Lhca1 binds to the core in higher plants, while the other is located in

a position that is not present in higher plants (Fig. 11B,C). Regarding the other Lhca

complexes, the improved map allows recognition of small, but strong, circular densities

Page 60: University of Groningen From Photosystem I to Photosystem

59

(Fig. 11A,B, black asterisks) that we assume correspond to helix C of the Lhca proteins,

which is roughly perpendicular to the membrane plane, thus resulting in a strong density in

the projection map (Fig. 10A). Based on the X-ray model fitting, the inner row of Lhca

proteins in C. reinhardtii comprises four Lhca proteins, which can directly associate with the

core, in principle, in a mode similar to that of plants. However, a local mismatch in the X-ray

model fitting indicates that at least one Lhca protein of the inner row, the first from the top,

has a slightly different binding position/orientation in C. reinhardtii (Fig. 11A,B, yellow

arrows). The fitting of the plant PSI-LHCI structure into the C.r.PSI-LHCI map highlights an

additional area of the projection map, comprising an outer row of Lhca proteins absent in

the plant structure (Fig. 11A). The localization of the topmost Lhca in the outer row is

evident from the well-resolved EM density (Fig. 11A,B, blue arrows). The fitting of the four

remaining Lhca proteins is facilitated by the localization of their C helices (Fig. 11B), allowing

determination of the position of all 9 Lhca complexes in the structure of C.r.PSI-LHCI.

Figure 11. Structural assignment of the smaller (A) and larger (B) type of the C.r.PSI-LHCI complex by fitting the high-resolution structure of the plant PSI-LHCI complex (2WSC) (Amunts, Toporik et al.2010). The white arrow indicates the position of two peripheral subunits PsaL and PsaH. The yellow arrow shows the best-resolved density of Lhca protein of the inner row, which corresponds to the position of Lhca1 in the plant PSI-LHCI complex; the blue arrow indicates the density of the best-resolved Lhca protein of the outer row of LHCI. Black asterisks (A) mark the high-density spots of the proposed outer LHCI half-ring, which are presumed to correspond to the C helix of the Lhca polypeptide. A tentative fitting of the outer row of LHCI using the structure of the Lhca2 protein is indicated in panel B. Panel C: The PSI-LHCI-LHCII supercomplex of A. thaliana (adapted from Kouřil et al. (2005)) is shown at the same scale, for a direct comparison of the PSI antenna size between green algae and plants.

4. Discussion

In contrast to higher plants, a consistent picture of the composition and organization of the

antenna subunits in PSI of C. reinhardtii is not yet available. There is also no consensus

about the number of Lhca copies associated with PSI, since Lhca/core stoichiometries

Page 61: University of Groningen From Photosystem I to Photosystem

60

ranging from 4 to 14 have been proposed (Bassi, Soen et al. 1992; Germano, Yakushevska et

al. 2002; Kargul, Nield et al. 2003; Takahashi, Yasui et al. 2004; Kargul, Turkina et al. 2005;

Subramanyam, Jolley et al. 2006; Stauber, Busch et al. 2009). In this work we have purified a

His6-tagged version of C.r.PSI-LHCI after mild detergent solubilization of the membranes.

PSI-LHCI subcomplexes with different antenna sizes have been obtained after further

solubilization of the purified complex, suggesting that the association of at least part of the

antenna to the core is relatively loose, which could, at least partially, explain the

discrepancy in the number of Lhcas that have been found associated with the core.

4.1 Lhca/core stochiometry: 9 Lhca complexes are associated with PSI core

The largest PSI-LHCI complex purified was shown to be highly homogeneous (Fig. 7). Several

lines of evidence suggest that the Lhca/core ratio of this complex is 9: (i) the comparison of

the C.r.PSI-LHCI projections with either the A. thaliana complex or the pea PSI-LHCI crystal

structure suggests the presence of 5 extra Lhcas in the former (Fig. 11); (ii) the apparent

MW of C.r.PSI-LHCI is 770KDa, corresponding to a complex with 9 Lhcas; (iii) the Chl a/b

ratio of 4.4 of C.r.PSI-LHCI is consistent with the presence of 9 Lhca complexes with an

average Chl a/b ratio of 2.2, in agreement with the results on reconstituted Lhcas (Mozzo,

Mantelli et al. 2010); (iv) a PSI subcomplex with MW of 690 kDa (i.e. still far larger than the

plant supercomplex) was shown (by MS and EM) to lack two Lhca subunits and some core

subunits, compared to the largest plant PSI-LHCI supercomplex.

EM analysis shows that the C.r.PSI-LHCI preparation contains only two types of particles,

differing by some density at the PsaH/PsaL site. This density, however, is too small to

accommodate a Lhca complex. The preparation is thus homogeneous with respect to the

antenna composition. Comparison of the EM maps of C.r.PSI-LHCI and the A2 subcomplex

reveals the positions of Lhca2 and Lhca9 as two well-defined densities. These densities are

present in all particles in the PSI-LHCI fraction, implying a 1:1 stoichiometry of each

polypeptide with the core.

In a recent study, it was found that most of the Lhca complexes, including Lhca2 and Lhca9,

are present in substoichiometric amount respect to the core (Stauber, Busch et al. 2009).

However, these data were obtained by averaging the results of three preparations with

different Chl a/b ratios, and thus very likely with different antenna sizes. Indeed, Stauber et

al. (2009) suggest that 7.5 ± 1.4 Lhca proteins are present per core complex, with the

Page 62: University of Groningen From Photosystem I to Photosystem

61

highest value in that range being in good agreement with our results. Moreover, the Lhca

complexes that were suggested to be in a 1:1 ratio with the core are Lhca1, Lhca3, Lhca4

and Lhca7, which correspond (with the exception of Lhca4) to the complexes that are still

present in the smaller particles (Fig. 5A), and are thus likely to be more stably associated

with the core.

4.2 Disassembly of the complex

The sequential disassembly of the C.r.PSI-LHCI complex after secondary solubilization

provides information about the hierarchic organization of the complex. The first step in the

disassembly process (i.e. production of subcomplex A2) corresponds to the loss of PsaL,

PsaH, Lhca2 and Lhca9. The fact that these polypeptides are lost first in the presence of

a very low concentration of ionic detergent indicates that they are rather loosely bound.

It also explains previous reports in which Lhca2 and Lhca9 were either absent or present in

substoichiometric amounts (Takahashi, Yasui et al. 2004; Stauber, Busch et al. 2009).

The facile loss of these LHCI complexes, which are known to harbour the red-most forms of

the antenna pigments (Mozzo, Mantelli et al. 2010), might also explain the large

discrepancies in excitation energy transfer and trapping studies of C.r.PSI-LHCI found in the

literature (Owens, Webb et al. 1988; Owens, Webb et al. 1989; Melkozernov, Kargul et al.

2004; Ihalainen, van Stokkum et al. 2005; Melkozernov, Kargul et al. 2005).

The second disassembly step leads to a complex (A3) with an apparent MW of 570 kDa, and

which lacks Lhca 4, Lhca5 and Lhca6. No loss of core subunits beyond PsaH and PsaL, which

are already absent in the A2 complex, was observed. The 120-kDa difference between A2

and A3 can thus be fully attributed to the loss of 3 Lhca antenna complexes.

The next subcomplex (A4) has an apparent molecular weight of 460 kDa. It has lost PsaK,

PsaG and PsaF, but it seems to retain part of the antenna. Indeed, all Lhcas but Lhca2 and

Lhca9 could be detected in this fraction. However, their level was around 40% of that in the

largest complex (A1), with the exception of Lhca7 and Lhca8, whose presence was

unchanged. Fraction A4 also shows a lower Chl a/b ratio as compared with fraction A3, in

agreement with a higher content of Lhca. The apparent discrepancy of having a complex

with reduced MW but with higher antenna content can be reconciled by taking into account

the fact that Lhca complexes can also form a large complex (300-540kDa) in the absence of

the core (Takahashi, Yasui et al. 2004; Ozawa, Onishi et al. 2010). This complex has been

Page 63: University of Groningen From Photosystem I to Photosystem

62

shown to be composed of all Lhcas but Lhca3, Lhca2 and Lhca9, and to have an emission

maximum at 708 nm. These are also the characteristics of the A4 fraction (except for the

presence of Lhca3), suggesting that this fraction contains a PSI subcomplex co-migrating

with a Lhca-oligomer. The 460-kDa complex corresponds well to a core that has lost PsaK,

PsaG, PsaF, PsaL and PsaH, but still retains two Lhcas. The best candidates for these are

Lhca7 and Lhca8, which are still present at WT levels, or Lhca3, which is present in this

fraction but it is not expected to be part of the Lhca-oligomer.

Fraction A5, with an apparent MW of 370 kDa, represents a core complex that has lost the

full Lhca complement, in addition to the above mentioned PSI-core subunits. The Lhca

proteins detected in this fraction (all but Lhca2 and 9) are all present at low but comparable

levels. This is likely due to co-migration of an antenna oligomer, which can apparently

exhibit different sizes.

In the last step of the disassembly, an additional complex appears, which has lost the

stromal subunits PsaC, PsaD and PsaE. It is composed of only the central subunits PsaA and

PsaB, and thus represents a stripped-down PsaA/B heterodimer. The implied disassembly

sequence: PsaH/L > PsaG/K/F > PsaC/D/E > PsaA/B suggests that the assembly of the PSI

complex occurs in the opposite order, in partial agreement with recent results (Ozawa,

Onishi et al. 2010). Moreover, it corresponds to more and more fundamental biochemical

function, with the PsaA/B heterodimer being the locus of charge separation, PsaC/D/E

adding the terminal acceptors and ferredoxin-binding site, PsaF/J (the latter implied) adding

the plastocyanin-binding site and PsaG/K adding the ability to bind LHCI antenna complexes

(Fischer, Hippler et al. 1998; Fischer, Boudreau et al. 1999; Scheller, Jensen et al. 2001;

Jensen, Bassi et al. 2007). Finally, PsaL/H add the ability to bind LHCII trimers during state

transitions (Lunde, Jensen et al. 2000; Zhang and Scheller 2004).

4.3 Photosystem I organization

Although several studies have focused on the structural organization of the PSI-LHCI

complex of C. reinhardtii (Germano, Yakushevska et al. 2002; Kargul, Nield et al. 2003;

Kargul, Turkina et al. 2005; Busch and Hippler 2011), this has not yet led to a conclusive

model for the antenna organization. The discrepancy between previous models is largely

due to limited resolution in the electron microscopy density maps. This made it impossible

to determine the positions of individual Lhca proteins, and even the orientation

Page 64: University of Groningen From Photosystem I to Photosystem

63

and positioning of the PSI core moiety can be questioned (Kargul, Nield et al. 2003).

Moreover, at the time of the first studies (Germano, Yakushevska et al. 2002; Kargul, Nield

et al. 2003), the plant PSI-LHCI structure, which turned out to be crucial for assignment of

the correct stoichiometry and position of the four Lhca antenna proteins (Ben-Shem, Frolow

et al. 2003), was not yet available. In this study we obtained a PSI-LHCI projection map at

about 15 Å resolution and the density profile enabled us to position the core complex and

nine Lhca antenna polypeptides. In particular, the assignment of 5 similar high-density spots

in the peripheral part of the LHCI antenna, assumed to correspond to the density of the C

helix of the 5 outer Lhca proteins, was helpful to establish the stoichiometry. The positions

of Lhca2 and Lhca9 were assigned with high confidence. Further data are necessary to

determine the exact location of the other 7 Lhca proteins, but based on the reasonable

assumption that PSI-LHCI complexes disassemble by losing the most external antenna

complexes first, we propose that Lhca4, Lhca5 and Lhca6 are located in the outer half-ring of

the supercomplex, while Lhca1, Lhca3, Lhca7 and/or Lhca8 compose the inner half-ring, with

Lhca3 located next to PsaK, based on previous results (Naumann, Stauber et al. 2005).

It is interesting to observe that the loss of PsaG, PsaK and PsaF leads to the dissociation of

a large part of the antenna, confirming the important role of these subunits in the stable

association of LHCI with PSI (Ozawa, Onishi et al. 2010).

4.4 PSI of Chlamydomonas vs. PSI of higher plants

The comparison of the PSI-LHCI supercomplexes from green algae and plants reveals that

they differ in several respects.

- The outer antenna of green algal PSI-LHCI in is larger than in plants (9 Lhcas vs. 4)

- The Lhcas are organized into two half-rings on one side of the core; the outer half-ring is

not present in plant supercomplexes.

- C.r.PSI-LHCI is less stable than the complex from higher plants and easily loses some of

the Lhcas, starting with Lhca9 and Lhca2.

- The amount of red forms is far lower for C.r.PSI-LHCI than for the plant supercomplex;

the C. reinhardtii complex has an emission maximum at 716 nm vs. 735 nm for the one

from A. thaliana.

- LHCI complexes from C. reinhardtii have a higher Chl b content as compared with plant

LHCI.

Page 65: University of Groningen From Photosystem I to Photosystem

References

Amunts, A., O. Drory, et al. (2007). "The structure of a plant photosystem I supercomplex at 3.4 A resolution." Nature 447(7140): 58-63.

Amunts, A. and N. Nelson (2008). "Functional organization of a plant Photosystem I: evolution of a highly efficient photochemical machine." Plant Physiol Biochem 46(3): 228-237.

Amunts, A. and N. Nelson (2009). "Plant photosystem I design in the light of evolution." Structure 17(5): 637-650.

Amunts, A., H. Toporik, et al. (2010). "Structure determination and improved model of plant photosystem I." Journal of Biological Chemistry 285(5): 3478-3486.

Arnon, D. I. (1971). "The light reactions of photosynthesis." Proc Natl Acad Sci U S A 68(11): 2883-2892.

Bassi, R. (1985). "Spectral Properties and Polypeptide Composition of the Chlorophyll-Proteins from Thylakoids of Granal and Agranal Chloroplasts of Maize (Zea-Mays-L)." Carlsberg Research Communications 50(2): 127-143.

Bassi, R., S. Y. Soen, et al. (1992). "Characterization of chlorophyll a/b proteins of photosystem I from Chlamydomonas reinhardtii." Journal of Biological Chemistry 267(36): 25714-25721.

Ben-Shem, A., F. Frolow, et al. (2003). "Crystal structure of plant photosystem I." Nature 426(6967): 630-635.

Ben-Shem, A., F. Frolow, et al. (2004). "Light-harvesting features revealed by the structure of plant photosystem I." Photosynthesis Research 81(3): 239-250.

Boekema, E. J., J. P. Dekker, et al. (1987). "Evidence for a Trimeric Organization of the Photosystem-I Complex from the Thermophilic Cyanobacterium Synechococcus Sp." Febs Letters 217(2): 283-286.

Boekema, E. J., H. Van Roon, et al. (1999). "Supramolecular organization of photosystem II and its light-harvesting antenna in partially solubilized photosystem II membranes." European Journal of Biochemistry 266(2): 444-452.

Busch, A. and M. Hippler (2011). "The structure and function of eukaryotic photosystem I." Biochim Biophys Acta 1807(8): 864-877.

Croce, R., G. Canino, et al. (2002). "Chromophore organization in the higher-plant photosystem II antenna protein CP26." Biochemistry 41(23): 7334-7343.

Dekker, J. P. and E. J. Boekema (2005). "Supramolecular organization of thylakoid membrane proteins in green plants." Biochim Biophys Acta 1706(1-2): 12-39.

Elrad, D. and A. R. Grossman (2004). "A genome's-eye view of the light-harvesting polypeptides of Chlamydomonas reinhardtii." Curr Genet 45(2): 61-75.

Farah, J., F. Rappaport, et al. (1995). "Isolation of a psaF-deficient mutant of Chlamydomonas reinhardtii: efficient interaction of plastocyanin with the photosystem I reaction center is mediated by the PsaF subunit." Embo Journal 14(20): 4976-4984.

Fischer, N., E. Boudreau, et al. (1999). "A large fraction of PsaF is nonfunctional in photosystem I complexes lacking the PsaJ subunit." Biochemistry 38(17): 5546-5552.

Fischer, N., M. Hippler, et al. (1998). "The PsaC subunit of photosystem I provides an essential lysine residue for fast electron transfer to ferredoxin." Embo Journal 17(4): 849-858.

Fischer, N., P. Setif, et al. (1997). "Targeted mutations in the psaC gene of Chlamydomonas reinhardtii: preferential reduction of FB at low temperature is not accompanied by altered electron flow from photosystem I to ferredoxin." Biochemistry 36(1): 93-102.

Ganeteg, U., F. Klimmek, et al. (2004). "Lhca5--an LHC-type protein associated with photosystem I." Plant Mol Biol 54(5): 641-651.

Germano, M., A. E. Yakushevska, et al. (2002). "Supramolecular organization of photosystem I and light- harvesting complex I in Chlamydomonas reinhardtii." Febs Letters 525(1-3): 121-125.

Page 66: University of Groningen From Photosystem I to Photosystem

65

Gibasiewicz, K., A. Szrajner, et al. (2005). "Characterization of low-energy chlorophylls in the PSI-LHCI supercomplex from Chlamydomonas reinhardtii. A site-selective fluorescence study." Journal of Physical Chemistry B 109(44): 21180-21186.

Gorman, D. S. and R. P. Levine (1965). "Cytochrome f and plastocyanin: their sequence in the photosynthetic electron transport chain of Chlamydomonas reinhardi." Proc Natl Acad Sci U S A 54(6): 1665-1669.

Gulis, G., K. V. Narasimhulu, et al. (2008). "Purification of His6-tagged Photosystem I from Chlamydomonas reinhardtii." Photosynthesis Research 96(1): 51-60.

Haldrup, A., H. Naver, et al. (1999). "The interaction between plastocyanin and photosystem I is inefficient in transgenic Arabidopsis plants lacking the PSI-N subunit of photosystem I." Plant Journal 17(6): 689-698.

Ihalainen, J. A., M. Ratsep, et al. (2003). "Red spectral forms of chlorophylls in green plant PSI - a site-selective and high-pressure spectroscopy study." Journal of Physical Chemistry B 107(34): 9086-9093.

Ihalainen, J. A., I. H. van Stokkum, et al. (2005). "Kinetics of excitation trapping in intact Photosystem I of Chlamydomonas reinhardtii and Arabidopsis thaliana." Biochim Biophys Acta 1706(3): 267-275.

Jansson, S. (1999). "A guide to the Lhc genes and their relatives in Arabidopsis." Trends Plant Sci 4(6): 236-240.

Jensen, P. E., R. Bassi, et al. (2007). "Structure, function and regulation of plant photosystem I." Biochimica Et Biophysica Acta-Bioenergetics 1767(5): 335-352.

Jensen, P. E., M. Gilpin, et al. (2000). "The PSI-K subunit of photosystem I is involved in the interaction between light-harvesting complex I and the photosystem I reaction center core." Journal of Biological Chemistry 275(32): 24701-24708.

Jensen, P. E., A. Haldrup, et al. (2003). "Molecular dissection of photosystem I in higher plants: topology, structure and function." Physiologia Plantarum 119(3): 313-321.

Jordan, P., P. Fromme, et al. (2001). "Three-dimensional structure of cyanobacterial photosystem I at 2.5 A resolution." Nature 411(6840): 909-917.

Kargul, J., J. Nield, et al. (2003). "Three-dimensional reconstruction of a light-harvesting complex I-photosystem I (LHCI-PSI) supercomplex from the green alga Chlamydomonas reinhardtii. Insights into light harvesting for PSI." Journal of Biological Chemistry 278(18): 16135-16141.

Kargul, J., M. V. Turkina, et al. (2005). "Light-harvesting complex II protein CP29 binds to photosystem I of Chlamydomonas reinhardtii under State 2 conditions." Febs Journal 272(18): 4797-4806.

Kouril, R., A. Zygadlo, et al. (2005). "Structural characterization of a complex of photosystem I and light-harvesting complex II of Arabidopsis thaliana." Biochemistry 44(33): 10935-10940.

Lam, E., W. Oritz, et al. (1984). "Isolation and Characterization of a Light-Harvesting Chlorophyll a/b Protein Complex Associated with Photosystem I." Plant Physiology 74(3): 650-655.

Liu, H., R. G. Sadygov, et al. (2004). "A model for random sampling and estimation of relative protein abundance in shotgun proteomics." Anal Chem 76(14): 4193-4201.

Lucinski, R., V. H. Schmid, et al. (2006). "Lhca5 interaction with plant photosystem I." Febs Letters 580(27): 6485-6488.

Lunde, C., P. E. Jensen, et al. (2000). "The PSI-H subunit of photosystem I is essential for state transitions in plant photosynthesis." Nature 408(6812): 613-615.

Melkozernov, A. N., J. Kargul, et al. (2004). "Energy coupling in the PSI-LHCI supercomplex from the green alga Chlamydomonas reinhardtii." Journal of Physical Chemistry B 108(29): 10547-10555.

Melkozernov, A. N., J. Kargul, et al. (2005). "Spectral and kinetic analysis of the energy coupling in the PSI-LHC I supercomplex from the green alga Chlamydomonas reinhardtii at 77 K." Photosynthesis Research 86(1-2): 203-215.

Page 67: University of Groningen From Photosystem I to Photosystem

66

Mozzo, M., M. Mantelli, et al. (2010). "Functional analysis of Photosystem I light-harvesting complexes (Lhca) gene products of Chlamydomonas reinhardtii." Biochim Biophys Acta 1797(2): 212-221.

Naumann, B., E. J. Stauber, et al. (2005). "N-terminal processing of Lhca3 Is a key step in remodeling of the photosystem I-light-harvesting complex under iron deficiency in Chlamydomonas reinhardtii." Journal of Biological Chemistry 280(21): 20431-20441.

Nelson, N. and A. Ben-Shem (2005). "The structure of photosystem I and evolution of photosynthesis." Bioessays 27(9): 914-922.

Oostergetel, G. T., W. Keegstra, et al. (1998). "Automation of specimen selection and data acquisition for protein electron crystallography." Ultramicroscopy 74(1-2): 47-59.

Owens, T. G., S. P. Webb, et al. (1988). "Antenna Structure and Excitation Dynamics in Photosystem-I .1. Studies of Detergent-Isolated Photosystem-I Preparations Using Time-Resolved Fluorescence Analysis." Biophysical Journal 53(5): 733-745.

Owens, T. G., S. P. Webb, et al. (1989). "Antenna Structure and Excitation Dynamics in Photosystem-I .2. Studies with Mutants of Chlamydomonas-Reinhardtii Lacking Photosystem-Ii." Biophysical Journal 56(1): 95-106.

Ozawa, S., T. Onishi, et al. (2010). "Identification and characterization of an assembly intermediate subcomplex of photosystem I in the green alga Chlamydomonas reinhardtii." Journal of Biological Chemistry 285(26): 20072-20079.

Peter, G. F. and J. P. Thornber (1991). "Biochemical composition and organization of higher plant photosystem II light-harvesting pigment-proteins." Journal of Biological Chemistry 266(25): 16745-16754.

Pineau, B., C. Gerard-Hirne, et al. (2001). "Carotenoid binding to photosystems I and II of Chlamydomonas reinhardtii cells grown under weak light or exposed to intense light." Plant Physiology and Biochemistry 39(1): 73-85.

Romero, E., M. Mozzo, et al. (2009). "The Origin of the Low-Energy Form of Photosystem I Light-Harvesting Complex Lhca4: Mixing of the Lowest Exciton with a Charge-Transfer State." Biophysical Journal 96(5): L35-L37.

Scheller, H. V., P. E. Jensen, et al. (2001). "Role of subunits in eukaryotic Photosystem I." Biochimica Et Biophysica Acta-Bioenergetics 1507(1-3): 41-60.

Stauber, E. J., A. Busch, et al. (2009). "Proteotypic profiling of LHCI from Chlamydomonas reinhardtii provides new insights into structure and function of the complex." Proteomics 9(2): 398-408.

Stauber, E. J., A. Fink, et al. (2003). "Proteomics of Chlamydomonas reinhardtii light-harvesting proteins." Eukaryot Cell 2(5): 978-994.

Storf, S., S. Jansson, et al. (2005). "Pigment binding, fluorescence properties, and oligomerization behavior of Lhca5, a novel light-harvesting protein." Journal of Biological Chemistry 280(7): 5163-5168.

Subramanyam, R., C. Jolley, et al. (2006). "Characterization of a novel Photosystem I-LHCI supercomplex isolated from Chlamydomonas reinhardtii under anaerobic (State II) conditions." Febs Letters 580(1): 233-238.

Takahashi, Y., T. A. Yasui, et al. (2004). "Comparison of the subunit compositions of the PSI-LHCI supercomplex and the LHCI in the green alga Chlamydomonas reinhardtii." Biochemistry 43(24): 7816-7823.

Tokutsu, R., H. Teramoto, et al. (2004). "The light-harvesting complex of photosystem I in Chlamydomonas reinhardtii: protein composition, gene structures and phylogenic implications." Plant Cell Physiol 45(2): 138-145.

van Oort, B., A. Amunts, et al. (2008). "Picosecond fluorescence of intact and dissolved PSI-LHCI crystals." Biophysical Journal 95(12): 5851-5861.

Wientjes, E. and R. Croce (2011). "The light-harvesting complexes of higher-plant Photosystem I: Lhca1/4 and Lhca2/3 form two red-emitting heterodimers." Biochemical Journal 433(3): 477-485.

Page 68: University of Groningen From Photosystem I to Photosystem

67

Wientjes, E., G. T. Oostergetel, et al. (2009). "The role of Lhca complexes in the supramolecular organization of higher plant photosystem I." Journal of Biological Chemistry 284(12): 7803-7810.

Wollman, F. A. and P. Bennoun (1982). "A New Chlorophyll-Protein Complex Related to Photosystem-I in Chlamydomonas-Reinhardii." Biochimica Et Biophysica Acta 680(3): 352-360.

Zhang, S. P. and H. V. Scheller (2004). "Light-harvesting complex II binds to several small Subunits of photosystem I." Journal of Biological Chemistry 279(5): 3180-3187.

Page 69: University of Groningen From Photosystem I to Photosystem

68

Page 70: University of Groningen From Photosystem I to Photosystem

69

CHAPTER 3

Light-harvesting complex II (LHCII) and its supramolecular

organization in Chlamydomonas reinhardtii

Bartlomiej Drop1#, Mariam Webber-Birungi2#, Sathish Yadav N.K.2, Alicja Filipowicz-

Szymanska3, Fabrizia Fusetti3, Egbert J. Boekema2 and Roberta Croce1

1 Department of Physics and Astronomy, Faculty of Sciences, VU University Amsterdam, De Boelelaan 1081, 1081 HV Amsterdam, The Netherlands.

2 Department of Biophysical Chemistry, Groningen Biological Sciences and Biotechnology Institute, University of Groningen, Nijenborgh 7, 9747 AG Groningen, The Netherlands;

3 Department of Biochemistry, Groningen Biomolecular Sciences and Biotechnology Institute, Netherlands Proteomics Centre & Zernike Institute for Advanced Materials, University of Groningen, Nijenborgh 4, 9747 AG Groningen, The Netherlands;

# these authors contributed equally to this work

Based on Biochimica Et Biophysica Acta 2014, 1837(1): 63-72.

Page 71: University of Groningen From Photosystem I to Photosystem

70

Abstract

LHCII is the most abundant membrane protein on earth. It participates in the first

steps of photosynthesis by harvesting sunlight and transferring excitation energy to the core

complex. Here we have analyzed the LHCII complex of the green alga Chlamydomonas

reinhardtii and its association with the core of photosystem II (PSII) to form multiprotein

complexes.

Several PSII supercomplexes with different antenna size have been purified, the largest of

which contains three LHCII trimers (named S, M and N) per monomeric core. A projection

map at 13 Å resolution was obtained allowing the reconstruction of the 3D structure of the

supercomplex. The position and orientation of the S trimer are the same as in plants; trimer

M is rotated by 45° degrees and the additional trimer (named here LHCII-N), which is taking

the position occupied in plants by CP24, is directly associated with the core. The analysis of

supercomplexes with different antenna size suggests that LhcbM1, LhcbM2/7 and LhcbM3

are the major components of the trimers in the PSII supercomplex, while LhcbM5 is part of

the “extra” LHCII pool not directly associated with the supercomplex. It is also shown that

Chlamydomonas LHCII has a slightly lower chlorophyll a/b ratio than the complex from

plants and a blue shifted absorption spectrum. Finally the data indicate that there are

at least six LHCII trimers per dimeric core in the thylakoid membranes, meaning that the

antenna size of PSII of Chlamydomonas reinhardtii is larger than that of plants.

Page 72: University of Groningen From Photosystem I to Photosystem

71

1. Introduction

Photosystem II (PSII) is a multisubunit pigment-protein complex located in the thylakoid

membrane of cyanobacteria, algae and higher plants. It captures and converts light into

chemical energy, which is used to oxidize water and reduce plastoquinone in the light

reactions of photosynthesis (Goussias, Boussac et al. 2002; Croce and van Amerongen

2011). In Arabidopsis thaliana (A.t.) the largest isolated PSII supercomplex is dimeric and

has a molecular weight of 1400 kDa (Caffarri, Kouril et al. 2009). The monomer may be

composed of up to 40 different proteins, most of which are permanently part of the

complex, while others are expressed or associated with it in stress conditions or during

assembly and degradation (Shi, Hall et al. 2012). In higher plants and algae, PSII is composed

of two moieties: the core complex that contains all the cofactors of the electron transport

chain, and the outer antenna, which increases the light-harvesting capacity of the core

(Dekker and Boekema 2005).

The PSII core complex is highly conserved in all organisms performing oxygenic

photosynthesis (Nield, Kruse et al. 2000; Buchel and Kuhlbrandt 2005; Umena, Kawakami et

al. 2011). The core contains the complex of the reaction center (RC, composed of D1 and D2

and cytochrome b559), that generates the redox potential required to drive water splitting

(Dau, Zaharieva et al. 2012) and the Chlorophyll (Chl) a-binding antenna complexes CP43

and CP47 (Dekker and Boekema 2005; Dang, Zazubovich et al. 2008). On the lumenal side of

the core, several extrinsic proteins (PsbO, PsbQ, PsbP, PsbR in plants and green algae) form

the oxygen-evolving complex (OEC), which supports water oxidation (Ifuku, Ishihara et al.

2010; Dau, Zaharieva et al. 2012).

The genes encoding for the antennas of PSII of plants and green algae are members of the

Light-harvesting complex (Lhc) multigenic family, which also includes the antenna

complexes of PSI (Jansson 1999). These proteins show structural homology (Ben-Shem,

Frolow et al. 2003; Liu, Yan et al. 2004): each Lhc polypeptide has three transmembrane α-

helices and coordinates Chls a, Chls b, and different carotenoid molecules. In higher plants

LHCII, the most abundant light-harvesting complex, is composed of 3 gene products (Lhcb1-

3) organized as heterotrimers. Each LHCII apoporotein binds eight Chls a, six Chls b and four

carotenoids (Liu, Yan et al. 2004; Standfuss, van Scheltinga et al. 2005). The other

chlorophyll a/b-binding proteins, Lhcb4, Lhcb5, and Lhcb6, also known as CP29, CP26, and

CP24, exist as monomers and have different pigment composition (Croce, Canino et al.

Page 73: University of Groningen From Photosystem I to Photosystem

72

2002; Passarini, Wientjes et al. 2009; Pan, Li et al. 2011). The major function of the outer

antenna system is to capture light energy and transfer excitation energy to the RC (Croce

and van Amerongen 2011). However, in high light, when the absorbed energy exceeds the

photosynthetic capacity and can be damaging for the system, it ensures photoprotection by

the dissipation of excess energy as heat (via the process called non-photochemical

quenching) to avoid formation of harmful radicals, e.g. (Muller, Li et al. 2001; de Bianchi,

Ballottari et al. 2010).

In C. reinhardtii nine genes (LhcbM1-M9) encode LHCII proteins. Two minor antennas, CP26

and CP29, are present while CP24 is not found in the genome of this alga (Elrad and

Grossman 2004; Minagawa and Takahashi 2004; Merchant, Prochnik et al. 2007). The LhcbM

gene products have been divided in four groups based on their sequence similarity: Type I

(LhcbM3, LhcbM4, LhcbM6, LhcbM8, and LhcbM9), Type II (LhcbM5), Type III (LhcbM2 and

LhcbM7), and Type IV (LhcbM1) (Teramoto, Ono et al. 2001; Minagawa and Takahashi

2004). Within their types, these proteins share a very high homology, i.e. sequence analysis

of the Type I or Type III proteins shows up to 99% identity. Instead, the average homology

between the types is lower: sequence identity between LhcbM6 (Type I) and LhcbM5 (Type

II), LhcbM2 (type III), LhcbM1 (type IV) is 80%, 77%, and 74%, respectively (Minagawa and

Takahashi 2004). Functional studies of these complexes are available only for LhcbM2/7,

LhcbM5 and LhcbM1 (Elrad, Niyogi et al. 2002; Takahashi, Iwai et al. 2006; Ferrante,

Ballottari et al. 2012) and suggest that the first two are involved in state transitions, while

LhcbM1 is important for NPQ.

The structures of several PSII components, including the cyanobacteria core (Guskov,

Kern et al. 2009; Umena, Kawakami et al. 2011), the plant LHCII trimer (Liu, Yan et al. 2004;

Standfuss, van Scheltinga et al. 2005) and CP29 from plants (Pan, Li et al. 2011) have been

resolved at intermediate/high resolution. Information about the organization of the PSII-

LHCII supercomplexes has been obtained by electron microscopy and single particle analysis

(Boekema, van Roon et al. 1999; Nield, Kruse et al. 2000; Nield, Orlova et al. 2000;

Yakushevska, Jensen et al. 2001; Yakushevska, Keegstra et al. 2003; Caffarri, Kouril et al.

2009; Tokutsu, Kato et al. 2012) and at present the best map has a maximal resolution of 12

Å (Caffarri, Kouril et al. 2009). The LHCII trimers associated with the core can be

distinguished in three types, based on their position in the PSII supercomplex and their

strong (S), moderate (M) or loose (L) association with the core (C) (Boekema, van Roon et al.

Page 74: University of Groningen From Photosystem I to Photosystem

73

1999). The C2S2M2 supercomplex is the largest PSII-LHCII observed in Arabidopsis thaliana

and it is composed of a dimeric core (C2), two LHCII-S trimers, interacting directly with the

core and with CP29 and CP26, and two M-trimers, which are associated with the core via

CP29 and CP24 (Boekema, van Roon et al. 1999; Nield, Orlova et al. 2000; Yakushevska,

Keegstra et al. 2003; Dekker and Boekema 2005; Caffarri, Kouril et al. 2009). The position of

the loosely bound trimers LHCII-L is still unclear and complexes containing one L trimer have

been observed only in spinach (Boekema, van Roon et al. 1999).

In Chlamydomonas reinhardtii a complex with C2S2 organization was reported (Nield, Kruse

et al. 2000; Iwai, Takahashi et al. 2008), as expected due to the absence of CP24, that in

higher plants is essential for connecting trimer M to the core (Kovacs, Damkjaer et al. 2006;

de Bianchi, Dall'Osto et al. 2008). However, recently, larger PSII supercomplexes with up to

6 trimers per dimeric core were observed (Tokutsu, Kato et al. 2012), showing that the

absence of CP24 does not influence the association of timer M with the supercomplex, and

that an additional timer is associated with the supercomplex on the side that in higher

plants is occupied by CP24.

In order to obtain further information about the antenna complexes of PSII and their

supramolecular organization in Chlamydomonas reinhardtii we have isolated different PSII

sub- and supercomplexes. These complexes were characterized by combining biochemistry,

spectroscopy and single particle electron microscopy. The analysis reveals the structural

organization of the PSII-LHCII and addresses the position and the composition of individual

LHCII trimers and their role in the assembly and functioning of the supercomplex.

2. Material and methods

2.1. Strain and growth conditions

Cells of Chlamydomonas reinhardtii strain (JVD-1B[pGG1]) in which a hexa-histidine tag has

been added at the N-terminus of PsaA core subunit of PSI (Gulis, Narasimhulu et al. 2008)

were grown in liquid Tris-Acetate-Phosphate (TAP) medium (Gorman and Levine 1965)

at room temperature (25°C) on an incubator shaker (Minitron, INFORS HT) at 170 rpm under

a continuous illumination flux of 20 μmol photons PAR m-2 s-1. In those conditions the cells

were in state 1.

Page 75: University of Groningen From Photosystem I to Photosystem

74

2.2. Thylakoids preparations

Cells were harvested by centrifugation (3500 rpm, 5 min, 4°C) at mid-logarithmic phase

(OD750nm ≈ 0.7). C.r. thylakoid membranes were prepared under dim light in a cold room as

described in (Fischer, Setif et al. 1997) with modification from (Drop, Webber-Birungi et al.

2011). PSII enriched membranes (BBY) from Arabidopsis thaliana were prepared according

to (Berthold, Babcock et al. 1981) with the modifications reported in (Caffarri, Kouril et al.

2009).

2.3. Isolation of PSII light-harvesting antenna complexes

PSII supercomplexes isolation was modified from (Drop, Webber-Birungi et al. 2011).

Thylakoids were pelleted, unstacked with 5 mM EDTA and washed with 10 mM Hepes (pH

7.5). Membranes were then resuspended in 20 mM Hepes (pH 7.5), 0.15 M NaCl and

solubilized at a final chlorophyll concentration of 0.5 mg/ml by adding an equal volume of

0.6% α-dodecylmaltoside (α-DM). Unsolubilized material was eliminated by centrifugation

(12,000 rpm for 10 min at 4°C).

To remove His-tagged PSI complexes from the preparation, the supernatant was loaded

onto a HisTrap HP Column (GE Healthcare) equilibrated with 20 mM Hepes (pH 7.5), 0.15 M

NaCl, 0.03% α-DM. PSI-depleted, fraction (flow through) was loaded on a sucrose density

gradient (prepared by freezing and thawing 0.5 M sucrose, 20 mM Hepes (pH 7.5), 0.03% α-

DM buffer layered over 1ml of 2M sucrose). PSII complexes were separated by

ultracentrifugation (41000 rpm, 14 h, 4°C). The green bands visible on the sucrose gradient

were harvested with a syringe.

The purification of PSII supercomplexes from A.t. was performed in the same conditions

used for the purification of the C.r. PSII supercomplexes, with the difference that upon

solubilization the membranes (BBY) were directly loaded on the sucrose gradient.

2.4. Gel electrophoresis

Proteins were analyzed by a SDS-6M urea PAGE with Tris-sulphate buffer system prepared

as in (Bassi 1985) at 14% acrylamide concentration. The amounts of sample loaded into each

well were: 3 µg (in Chls) for thylakoids; 2.5 µg for PSII supercomplexes; 1.5 µg for Lhcb

fractions. The Coomassie stained gel was imaged with ImageQuant LAS-4000 (GE

Healthcare).

Page 76: University of Groningen From Photosystem I to Photosystem

75

2.5. Mass spectrometry analysis

2.5.1. In-gel tryptic digestion

For mass spectrometry-based protein identification, the SDS-PAGE bands were excised from

the gel and treated with 10 mM DTT followed by 55 mM iodoacetamide in 50 mM NH4HCO3,

to reduce and alkylate cysteine residues, and subsequently dehydrated by incubation for 5

min in 100% acetonitrile. The gel slices were rehydrated in 10 µl trypsin solution (Trypsin

Gold, Mass spectrometry grade, Promega10 ng/µl in 25 mM NH4HCO3), and incubated for 2

hours at 37°C. Subsequently 10 µl of 25 mM NH4HCO3 was added to prevent drying and the

incubation was prolonged overnight at 37°C. The tryptic peptides were recovered by three

subsequent extractions with 50 µl of 35%, 50% and 70% acetonitrile in 0.1% TFA.

The extracted peptides were pooled and concentrated under vacuum.

To determine protein composition of PSII supercomplexes, sucrose gradient fractions were

first loaded on 10% SDS-PAGE and run about 1 cm through the resolving gel. This procedure

was applied to clean up samples from detergent. The whole gel bands were then cut in

three slices and tryptic digestion was performed as described above.

2.5.2. Liquid Chromatography-Mass Spectrometry (LC-MS)

Fractions of the peptide mixtures from in-gel trypsin digestions were diluted in 5% formic

acid, passed through a pre-column (EASY-Column C18, 100 µm x 20mm, 5µm particle size,

Thermo Scientific, Bremen, Germany) and separated on a capillary column (C18 PepMap

300, 75 µm x 100 mm, 3-µm particle size, Thermo, Thermo Scientific, Bremen, Germany)

mounted on a Proxeon Easy-nLCII system (Thermo Scientific, Bremen, Germany). Solutions

of 0.1% formic acid in water and a 0.1% formic acid in 100% acetonitrile were used as the

mobile phases. A gradient from 4 to 35% acetonitrile was performed in 120 min at a flow

rate of 300 nL/min. Eluted peptides were analyzed using a Linear ion Trap-Orbitrap hybrid

mass spectrometer (LTQ-Orbitrap XL, Thermo Scientific). MS scans were acquired in the

Orbitrap, in the range from 350 to 1800 m/z, with a resolution of 60,000 (FWHM).

The 7 most intense ions per scan were submitted to MS/MS fragmentation (35% Normalized

Collision EnergyTM) and detected in the linear ion trap.

Page 77: University of Groningen From Photosystem I to Photosystem

76

2.5.3. Protein identification

The MS raw data were analysed with Mascot (version 2.1, Matrix Science, London, UK) using

the Proteome Discoverer 1.3 analysis platform (Thermo Scientific) and searched against the

C.r. proteome. Peptide tolerance was set to 10 ppm and 2.0 Da for intact peptides and

fragment ions respectively, using semi-trypsin as protease specificity and allowing for up to

2 missed cleavages. Oxidation of methionine residues, deamidation of asparagine and

glutamine, and carbamidomethylation of cysteines were specified as variable modifications.

The MS/MS based peptide and protein identifications were further validated with the

program Scaffold (version Scaffold 4.0, Proteome Software Inc., Portland, OR). Protein

identifications based on at least 2 unique peptides identified by MS/MS, each with

a confidence of identification probability higher than 95%, were accepted.

2.6. Pigment composition

The chlorophyll concentrations of thylakoid preparations were calculated in 80% (v/v)

acetone, according to (Porra, Thompson et al. 1989).

The pigment composition of the complexes was analyzed by fitting the spectrum of the 80%

acetone extracted pigments with the spectra of the individual pigments in acetone and by

HPLC, as described previously (Croce, Canino et al. 2002). As shown in (Angeler and Schagerl

1997), loroxanthin was eluted just after neoxanthin, but in our experimental setups the

separation of these two was not possible: both carotenoids resulted in single peak. The data

are the results of at least four different preparations in two replicas.

2.7. Spectroscopic analysis

Room temperature absorption spectra were recorded with a Cary 4000 spectrophotometer

(Varian).

The fluorescence emission spectra were recorded at 5°C at low temperature (77K) using

a Fluorolog 3.22 spectrofluorimeter (Jobin Yvon-Spex). For 77K measurements a home built

liquid nitrogen cooled device was used. The excitation wavelengths were 440 nm, 475 nm

and 500 nm and emission was detected in the 600-800 nm range. Excitation and emission

slits bandwidth were set to 3 nm. All fluorescence spectra were measured at OD 0.05 at the

maximum of the Qy absorption. Room temperature measurements were performed in

0.5 M sucrose, 20 mM Hepes (pH 7.5), 0.03% α-DM buffer. For low-temperature

Page 78: University of Groningen From Photosystem I to Photosystem

77

measurements, samples were in 66.7% glycerol (w/v), 20 mM Hepes (pH 7.5), 0.03 % α-DM

buffer.

2.8. Electrochromic shift (ECS)

PSI/PSII ratio of Chlamydomonas cells was measured with a Joliot-type spectrophotometer

(Joliot and Delosme 1974) (Bio-Logic SAS JTS-10) as described previously (Petroutsos,

Terauchi et al. 2009; Bailleul, Cardol et al. 2010). A PSI/PSII ratio of 0.97±0.2 was obtained.

2.9. Electron microscopy and single particle analysis

Samples were negatively stained using the droplet method with 2% uranyl acetate on glow

discharged carbon-coated copper grids. Electron microscopy was performed on a Philips

CM120 electron microscope equipped with a LaB6 filament operating at 120 kV. Images

were recorded with a Gatan 4000 SP 4K slow-scan CCD camera at either 130000 ×

magnification at a pixel size (after binning the images) of 2.25 Å, respectively, at the

specimen level with GRACE software for semi-automated specimen selection and data

acquisition (Oostergetel, Keegstra et al. 1998). Single particle analysis was performed using

GRIP software including multi-reference and non-reference alignments, multivariate

statistical analysis, and classification, as in (Boekema, van Roon et al. 1999).

To determine the angle of rotation of the trimers: 1) in GRIP, all three trimers were selected

at their midpoints, boxed out of the complex projections and aligned to their 3-fold

rotationally averaged projections before aligning with S-trimer as reference, and 2).

In comparison, PowerPoint (Microsoft) was used to draw a triangle as a reference, around

the reference trimer. Copies of this triangle were then rotated to align with the other

trimers and the angle by which the triangle was rotated was determined in the program.

To model the supramolecular organization of the supercomplex, the available crystal

structures of the cyanobacterial PSII core (Umena, Kawakami et al. 2011) (3ARC), LHCII

trimer (Liu, Yan et al. 2004) (1RWT) and CP29 (Pan, Li et al. 2011) (3PL9) were used. Pymol

was used to construct an aligned complex model.

Page 79: University of Groningen From Photosystem I to Photosystem

78

3. Results

3.1. Isolation of PSII supercomplexes

To purify the PSII-LHCII complexes of C. reinhardtii the procedure described before for

higher plants (Caffarri, Kouril et al. 2009) was modified. Since it is not possible to prepare

grana membranes from C. reinhardtii, as the presence of PSI interferes with the purification

of PSII supercomplexes, we have used a strain carrying a hexa-histidine tag at

the N-terminus of the core subunit PsaA (Gulis, Narasimhulu et al. 2008; Drop, Webber-

Birungi et al. 2011). The thylakoid membranes purified from this strain were mildly

solubilized, to keep the large PSII supercomplexes intact, and loaded on a Ni-Sepharose

column to eliminate PSI. To obtain homogeneous preparations of PSII sub- and super-

complexes with different antenna sizes, the flow through fraction was subjected to sucrose

density gradient ultracentrifugation. The gradient separation resulted in one yellow band

(B1, containing free pigments), eight green bands (B2–B9), containing protein-chlorophyll

complexes, and a fraction of partially unsolubilized material laying on top of the 2M sucrose

solution (B10) (Fig. 1A).

Figure 1. Purification of PSII sub- and supercomplexes. A: Fractionation of PSII sub/super-complexes by sucrose density gradient. B: SDS-PAGE analysis of fractions B2–B9 from sucrose gradient. Proteins identified by mass spectrometry are indicated.

The bands pattern was very similar to that obtained from the solubilization of the grana

membrane of A. thaliana, characterized before (Caffarri, Kouril et al. 2009), suggesting the

Page 80: University of Groningen From Photosystem I to Photosystem

79

presence of PSII sub- and supercomplexes. As expected, the band corresponding to the A.t.

CP24/CP29/LHCII-M complex (“band4”) was absent in C. reinhardtii, due to the lack of CP24.

The polypeptide composition of fractions B2-B9 was analyzed by SDS-PAGE (Fig. 1B),

showing the presence of PSII components in all fractions. The exact protein composition of

the most prominent gel bands was confirmed by mass spectrometry. The results of the SDS-

PAGE analysis are summarized in Fig. 1B. Fraction B2, which mobility in the gradient

corresponds to monomeric Lhcb, contains CP26 and CP29, and LHCII (mainly LhcbM1,

LhcbM2/M7 and LhcbM3), while fraction B3 contains the trimeric complexes. It was

suggested that the trimeric form of C.r.LHCII is less stable than that of plants and may partly

dissociate into monomers upon treatment with detergents (Bassi and Wollman 1991). This

can explain the high content of LHCII in fraction B2 and the higher abundance of monomeric

complexes (B2) compared to trimeric (B3) (Fig. 1AB). However, it should be mentioned that

when solubilized thylakoids were directly loaded on the sucrose gradient, the trimeric

fraction was far more abundant than that of monomers (data not shown) indicating that,

although less stable than in plants, almost of the LhcbMs are present as trimers in the

membrane. The LhcbM protein composition of B2 and B3 was very similar with the

exception of LhcbM5, which was only present in B3, suggesting that trimers containing

LhcbM5 are relatively stable. Interestingly, CP26, which in plants exists as a monomer, could

be observed mainly in fraction B3. There are two possible explanations: CP26 forms homo or

hetero-trimers in C. reinhardtii, or it is strongly associated with a LHCII trimer.

The polypeptide composition of fraction B4 revealed the presence of PSII core subunits but

also of LHCII and CP29. It is clear that the antenna present in this band cannot be associated

with the core because the molecular weight of this band corresponds to that of PSII core

monomer. This suggests that the monomeric core co-migrates with an oligomer of antenna

complexes. In A. thaliana a complex composed of CP24/CP29/LHCII-M could be purified.

In C. reinhardtii CP24 is absent but its position in the supercomplex is occupied by a LHCII

trimer (Tokutsu, Kato et al. 2012)(see below). It might thus be possible that a complex

composed of CP29 and 2 LHCII trimers is stable enough to survive solubilization. The

molecular weight of this complex would indeed be comparable to that of the monomeric

core. In contrast to B4, in fraction B5 the amount of CP26 increased significantly. PSII

subcomplex in B5 might consist of PSII monomeric core/LHCII-S/CP26 (CS/CP26), as was

reported before for higher plants (Caffarri, Kouril et al. 2009). The SDS-PAGE of fractions B4

Page 81: University of Groningen From Photosystem I to Photosystem

80

and B6 showed the presence of LHCI antenna indicating that these fractions are

contaminated with PSI sub- and super-complexes, which are indeed expected to migrate at

these positions (Drop, Webber-Birungi et al. 2011). The presence of PSI contamination is

confirmed by the spectra which are red-shifted compared to the spectra of the other

fractions (data not shown). In the upper part of the gel the bands of the α and β subunits of

the ATPase are also visible. Due to their heterogeneous content bands B4-B6 were thus not

analyzed further. The protein composition of B7, B8 and B9, which correspond to PSII

supercomplexes, was very similar, while the ratio between PSII core subunits (CP47 and/or

CP43) and Lhcb subunits decreased when going from B7 to B9, in agreement with an

increased antenna size.

3.2. Electron microscopy and single-particle analysis

To determine the structural organization of PSII-LHCII, electron microscopy and single

particle analysis were performed. The analysis of about 50,000 projections from fractions

B7, B8 and B9 yielded six types of supercomplexes (Fig. 2AB), in which the dimeric PSII core

complex (C2) was associated with a variable number of LHCII trimers.

Figure 2. A: EM analysis of the C.r. PSII supercomplexes obtained from fractions B7–9. Averaged projection maps of PSII particles obtained by single particle averaging of about 50,000 projections. (A) C2S2M2N2 particle, sum of 128 projections. (B) C2S2MN particle, sum of 5000 projections. (C) C2SMN particle, sum of 2000 projections. (D) C2S2 particle, sum of 1024 particles. (E) C2S, sum of 512 (F) C2 core sum of 128. Notes: the map of frame B has been high-pass filtered to enhance the fine features of image processing. A two-fold rotational symmetry was imposed after processing on the map of frame A. Spacebar for all frames equals 10 nm. B: Numbers of projections analyzed from fractions B7, B8 and B9.

Page 82: University of Groningen From Photosystem I to Photosystem

81

Two LHCII trimers occupied positions equivalent to those of the S- and M- trimers in the

C2S2M2 supercomplex of A. thaliana (Caffarri, Kouril et al. 2009), while the third trimer was

located in the position that in A. thaliana is occupied by CP24, as observed before (Tokutsu,

Kato et al. 2012). Because this trimer was directly associated with the PSII core, without

involvement of any monomeric antenna, it is named trimer-N (naked). We prefer this

notation to “trimer L” for two reasons: (i) trimer L has been observed only in spinach where

it occupies a different position; (ii) “L” is short for “loosely bound” (Boekema, van Roon et

al. 1999), which is not completely appropriate for this trimer because its association to the

supercomplex survives purification.

Fractions B7 and B8 contain mainly PSII dimeric core complexes with one S trimer (C2S) and

two S trimers (C2S2) respectively. Fraction B9 contains three types of PSII-LHCII particles,

which beside LHCII-S and LHCII-M also include LHCII-N trimers: C2SMN, C2S2MN and the

largest complex composed of 6 LHCII trimers surrounding the dimeric core, C2S2M2N2.

If we consider all the particles from the data set, and not only those added in the final

figures because of their good image quality, the ratio of the three largest particles in B9 was

about 4.5(C2SMN):4 (C2S2MN):1(C2S2M2N2). This means that in the lower band the average

number of trimers per dimeric core complex is about 3.7.

The projection of the C2S2MN supercomplex was obtained at 13 Å resolution. This resolution

allows to identify the major structural features of the individual complexes and to use them

to fit the X-ray high-resolution structures of LHCII (Liu, Yan et al. 2004), CP29 (Pan, Li et al.

2011) and PSII core (Umena, Kawakami et al. 2011) into the projection map of the particle.

The obtained model of the supramolecular organization of the PSII supercomplex is shown

in Fig. 3 and 4.

The projection maps of the C2S2 supercomplexes of C. reinhardtii and A. thaliana are very

similar, suggesting that CP26 and CP29 occupy the same positions, with CP26 located close

to CP43 and CP29 next to CP47 (Boekema, van Roon et al. 1999; Yakushevska, Keegstra et al.

2003; Caffarri, Kouril et al. 2009). Trimer S is also located in the same position and it has the

same orientation with respect to the core in the two supercomplexes. This is not the case

for trimer M, which in C. reinhardtii is rotated of 45° compared to its orientation in A.

thaliana. N-trimer is associated with the core, CP29 and trimer M and it is rotated of 25°

with reference to trimer S (Fig. 4B). These values are within an accuracy of 5-10° because

the trimer features are not precisely outlined by the uranyl acetate negative stain.

Page 83: University of Groningen From Photosystem I to Photosystem

82

This contrasting agent does not penetrate much inside the membrane, where the bulk of

the protein is located.

Figure 3. Map of the C.r.PSII-LHCII supercomplex at 13Å resolution. A: Top view projection map of the C2S2MN supercomplex obtained from single particle electron microscopy. B: Assignment of the subunits in the supercomplex by fitting the high-resolution structures of PSII core (Umena, Kawakami et al. 2011) (in lime green), trimeric LHCII-S and LHCII-M (in brown), novel LHCII-N (in red) and monomeric Lhcb (in magenta) (Liu, Yan et al. 2004), and CP29 (Pan, Li et al. 2011). C: The largest PSII-LHCII supercomplex (C2S2M2) isolated from A.t. (adapted from (Caffarri, Kouril et al. 2009)) is shown at the same scale for a direct comparison of the PSII antenna size between green algae and higher plants. Spacebar equals 10 nm.

Figure 4. A: Model of the structure of the C.r.PSII-LHCII supercomplex. The model has been assembled based on the projection map in Fig. 2 using the crystal structures of the cyanobacterial PSII core (Umena, Kawakami et al. 2011) (3ARC), LHCII trimer (Liu, Yan et al. 2004) (1RWT) and CP29 (Pan, Li et al. 2011) (3PL9). For CP26, the structure of a monomeric LHCII has been used. Proteins of the PSII core (lime green), LHCII-S and -M (brown), novel LHCII-N (red), CP29 and CP26 (magenta), Chls a (cyan), Chls b (green), Neoxanthin (yellow spheres), Lutein L1 (orange), Lutein L2 (dark-yellow sticks). B: Presentation of LHCII-S, LHCII-M and LHCII-N trimers selected from the C.r.C2S2MN PSII supercomplex to determine their rotated orientations. S-trimer orientation was used as a reference.

Page 84: University of Groningen From Photosystem I to Photosystem

83

3.3. Protein composition of the supercomplexes

To get information about the protein composition of the PSII supercomplexes, fractions B7-

B9 were analysed by mass spectrometry. A shotgun proteomics approach revealed the

presence of all major core subunits (D1, D2, CP43, CP47) as well as of the oxygen evolving

complex subunits (PsbP, PsbO, PsbR) and of some of the small core subunits (PsbE, PsbF,

PsbH, PsbW). The other core subunits are likely to be present in PSII supercomplexes, but

unidentifiable by tryptic digestion, because the resulting peptides would be too short and

would not be detectable in our experimental setup. The analysis also showed the presence

of CP26 (Lhcb5), CP29 (Lhcb4) and of several LhcbM gene products. The sequences of LhcbM

proteins are highly homologous, and the analysis allows to identify with 100% probability (at

least 2 exclusive peptides with >95% probability) LhcbM1, LhcbM2/7 (LhcbM2 and M7 are

practically identical and it is not possible to discriminate between the two), LhcbM3 and

LhcbM5, while only one peptide was identified for LhcbM8 and LhcbM9. However, several

peptides that are common to LhcbM4 and M6 were also detected, indicating that at least

one of these two proteins is present in the preparation.

Figure 5. Protein composition of PSII-LHCII supercomplexes. The number of spectra assigned to individual PSII core subunits and Lhcb monomers (CP26 and CP29) was normalized to the average number of spectra assigned to the core complex polypeptides (calculated as (D1+D2)/2) in the same fraction. The values are reported relatively to the values determined for each polypeptide in fraction B9.

To determine the relative abundance of each subunit in the different fractions, spectral

count normalization was applied, meaning that the spectra of each subunits were counted

and were normalized to the spectra of D1 and D2 in the same fraction to compensate for

differences in sample loading (Liu, Sadygov et al. 2004). The PSII core subunits content of

Page 85: University of Groningen From Photosystem I to Photosystem

84

B7-B9 is presented in Fig. 5 normalized to the abundance of the same subunit in B9. The

data indicate that the amount of the major core subunits - D1, D2, CP43 and CP47 -

remained almost unchanged, supporting the reliability of the method. Lhcb4 and Lhcb5

were present in the same amount in B8 and B9 while their content was reduced in B7, in

agreement with the presence of C2S complexes in this fraction. Because of the high

sequence identity between the LhcbM proteins, it was not possible to use spectral counts to

determine their relative amount in the three samples. However, the number of unique

spectra matched to the different LhcbM types is very different and implies that LhcbM2/7,

LhcbM1 and LhcbM3 are the most abundant LhcbM proteins in this alga.

3.4. Pigment composition

The pigment composition of fractions B2,B3 and B7–B9, was analyzed together with the

composition of C.reinhardtii cells. The results are reported in Table 1. A Chl a/b ratio of 1.29

and 1.28 was obtained for the fractions of monomeric (B2) and trimeric (B3) antenna

complexes, similar to what previously observed (Pineau, Gerard-Hirne et al. 2001) slightly

lower than the value of plant LHCII (1.33).

Table 1. Pigment composition of: LHCII trimers (B3) and C.r.PSII supercomplexes (B7-B9).The

values of individual carotenoids are normalized to 14 Chls (a+b) in fractions B3 and to 100

Chls (a+b) in fractions with PSII supercomplexes (B7-B9) and thylakoids. The data are the

results of at least four different preparations in two replicas.

Chl a/Chl b Chls/Cars Neoxanthin/

Lohroxanthin

Violaxanthin Lutein β-caroten

B2 1.29 ± 0.06 3.79 ± 0.54 2.41 ± 0.20 0.55 ± 0.07 0.91 ± 0.04 -

B3 (LHCII trimer) 1.28 ± 0.02 3.65 ± 0.11 1.67 ± 0.22 0.68 ± 0.05 1.44 ± 0.18 -

B7 2.94 ± 0.18 5.60 ± 0.39 5.75 ± 0.48 1.01 ± 0.09 5.42 ± 0.47 5.65 ± 0.50

B8 2.41 ± 0.17 4.80 ± 0.75 6.98 ± 0.93 1.03 ± 0.09 6.43 ± 0.91 6.40 ± 1.04

B9 2.22 ± 0.13 4.67 ± 0.38 7.09 ± 0.42 1.12 ± 0.11 8.57 ± 1.44 7.11 ± 1.92

C.r. cells 2.28 ± 0.061 4.51 ± 0.421 - - - -

Page 86: University of Groningen From Photosystem I to Photosystem

85

The trimer coordinates four carotenoids per 14 Chls as it is the case in plants.

The carotenoid composition of C. reinhardtii is considered to be similar to that of higher

plants (Bassi, Pineau et al. 1993) with the addition of loroxanthin, which derives from lutein

by hydroxylation of the methyl group at C9 of the polyene chain and is selectively bound to

LHCII (Knoetzel, Braumann et al. 1988; Grossman, Lohr et al. 2004). Indeed in the trimeric

fraction we observed the presence of loroxanthin and a lower content of lutein as compared

to higher plants LHCII (Bassi, Pineau et al. 1993). Although it was not possible to fully

separate neoxanthin from loroxanthin, the fact that all LhcbMs contain the tyrosine

responsible for the selectivity of the N1 binding site for neoxanthin (Caffarri, Passarini et al.

2007), suggests the binding of 1 neoxanthin per complex and thus of around 0.7 loroxanthin

per monomer. Loroxanthin is probably substituting lutein in one of the two internal binding

sites (L1 and L2).

3.5. Spectroscopic characterization

The absorption spectra of fractions B2-B9 were measured at room temperature and are

presented in Fig. 6. The spectra of fractions B2 and B3 (Fig. 6A) show maxima at 672 nm and

at 670 nm respectively. Interestingly the absorption maximum of the Cr. LHCII trimer is 4 nm

blue shifted compared to that of the complex of higher plants (Fig. 6B) and it shows

relatively more intense absorption around 650–653 nm, in agreement with the lower Chl

a/b ratio. The absorption spectra of fractions B7-B9 (Fig. 6C) showed maxima at 675 nm.

A relative increase in intensity of the absorption in the Chl b region (630-660 nm) was

observed going from B7 to B9 (Fig. 6C), in agreement with the increased antenna size of the

supercomplexes.

To check the integrity of PSII supercomplexes, fluorescence spectra from fractions B7-B9

were measured at 5°C (Fig. 7A). The maximum emission for all fractions was at 678 nm.

The perfect overlapping of the emission spectra after excitation at 440 nm, 475 nm and

500 nm, which excite preferentially Chl a, Chl b and carotenoids, indicates that there are no

free pigments in these preparations. In Fig. 7A only the emission spectra of B9 are shown,

but the results were very similar for all fractions.

Low temperature (77K) fluorescence emission spectra showed a major peak at 676 nm in all

three fractions and a shoulder at 690-700 nm which is likely due to the core complex

Page 87: University of Groningen From Photosystem I to Photosystem

86

(Fig. 7B). Moreover, the spectrum of the supercomplexes is clearly red-shifted as compared

to that of LHCII trimer, confirming that there is energy transfer between LHCII and the core.

Figure 6. Absorption spectra at room temperature of the fractions from sucrose gradient. The spectra were normalized to the maximum absorption of the Qy region. A: Absorption spectra of fractions B2 and B3. B: Comparison of the absorption spectra of C.r.LHCII and A.t.LHCII trimers. C: Absorption spectra of fractions B7-B9.

Figure 7. Fluorescence emission spectra of the PSII-LHCII supercomplexes. A: Fluorescence emission spectra of PSII-LHCIIsupercomplex from B9 fraction recorded at 5°C. B: Low temperature (77K) fluorescence emission spectra of PSII-LHCII supercomplexes (B8-B9) and C.r.LHCII trimers (B3) excited at 440 nm.

4. Discussion

4.1. In the C. reinhardtii PSII supercomplex LHCII-N trimer substitutes CP24 and stabilizes

the binding of trimer M (and the other way around).

Early work comparing a PSII supercomplex isolated from C. reinhardtii with the C2S2

supercomplex of spinach showed high similarity in shape and size (Nield, Kruse et al. 2000).

This, together with the fact that CP24, which is essential for the binding of trimer M in

higher plants (Kovacs, Damkjaer et al. 2006; de Bianchi, Dall'Osto et al. 2008; Caffarri, Kouril

et al. 2009), is absent in C. reinhardtii led to the conclusion that the PSII supercomplex of

Page 88: University of Groningen From Photosystem I to Photosystem

87

this alga is organized as C2S2. Recently, larger complexes were observed containing a novel

trimer (here called N) in addition to trimers S and M (Tokutsu, Kato et al. 2012). In our

study, the EM analysis of isolated C.r.PSII supercomplexes (B7-B9) showed that they were

present in C2S2M2N2, C2S2MN, C2SMN and C2S2 configurations, while differently from higher

plants, no C2S2M, C2SM, C2S2M2 or C2M (Caffarri, Kouril et al. 2009) could be found.

Apparently, the lack of the neighboring N-trimer strongly influences the binding of LHCII-M,

as does CP24 in plants. Indeed in higher plants a complex composed of CP24/CP29/LHCII

could be isolated (Dainese and Bassi 1991), indicating that the association between these

proteins is stronger than between them and the core. The CP24/CP29/LHCII complex is

obviously not present in C. reinhardtii, but the analysis of the sucrose gradient bands

indicates the presence of a CP29/LHCII-M/LHCII-N complex. This suggests that the

connection of trimers N and M is stronger than their association with the core and might

also explain why, after detergent solubilization, only C2S2 could be detected in early works

(Nield, Kruse et al. 2000; Iwai, Takahashi et al. 2008).

These results point out that the hypothesis that CP24 has evolved in land plants to increase

the antenna size of PSII (or to avoid the formation of C2S2) should be reconsidered as clearly

PSII-LHCII of C. reinhardtii is larger than the largest supercomplex of plants (Kovacs,

Damkjaer et al. 2006; de Bianchi, Dall'Osto et al. 2008). It has been shown that under high

light stress CP24 and LHCII-M dissociate from the supercomplex (Betterle, Ballottari et al.

2009). It is thus likely that CP24 (and Lhcb3) has evolved not to make the antenna size of PSII

larger, but more flexible, allowing for a fast regulation of the PSII antenna size.

4.2. LhcbM1, LhcbM2/7 and LhcbM3 are the main components of the PSII supercomplexes

while LhcbM5 is enriched in the “extra” LHCII population.

LHCII antenna in C. reinhardti is codified by 9 different genes, LhcbM1-M9, but very little is

known about their organization and occurrence. The possibility to compare the composition

of supercomplexes containing different trimers (S for B7 and B8 and S, M and N for B9) with

that of the LHCII trimers that were detached from the supercomplexes (B3) allows to

analyse the protein composition of each trimer. Four main LhcbM bands could be resolved

in SDS-PAGE, which were identified as the four types of LhcbM complexes. It is interesting to

observe that, while the ratio LhcbM/core is increasing when going from B7 to B9,

the relative abundance of the bands associated with type I, IV and III polypeptides remains

Page 89: University of Groningen From Photosystem I to Photosystem

88

the same. This suggests that all trimers, have similar protein composition where LhcbM1,

LhcbM2/7 and LhcbM3 are the dominant protein, as confirmed by mass spectrometry

identification. Interestingly the SDS-PAGE shows that LhcbM5 is highly enriched in B3, while

very little LhcbM5 is present in the supercomplexes, suggesting that it is mainly present in

the LHCII pool that is not directly associated with the core and that in plants we have

defined as “extra” LHCII (Wientjes, van Amerongen et al. 2013). In addition, LhcbM5,

together with LhcbM1, is the only protein that conserves the phosphorylation site which is

important for state transitions (Elrad and Grossman 2004) and it has indeed been proposed

to participate in this process (Stauber, Fink et al. 2003; Takahashi, Iwai et al. 2006).

4.3. The properties of C.r. LHCII differ from those of plants LHCII

Sequence analysis has shown that the LhcbM proteins of C. reinhardtii do not exactly

correlate with the Lhcb1-3 proteins of higher plants (Teramoto, Ono et al. 2001; Elrad and

Grossman 2004) although they show a high degree of identity. All amino acids which were

shown to be responsible for the Chl binding in plants (Liu, Yan et al. 2004) are conserved in

the nine LhcbM proteins suggesting that they bind the same number of Chls as

the complexes from plants. The Chl a/b ratio of the trimer differs only very slightly from that

of plants (1.28 in C. reinhardtii vs. 1.33 in plants) indicating at most the change of affinity of

one binding site. However, the absorption spectrum of C.r.LHCII trimer is different from that

of A.t.LHCII (Fig 6B) showing a blue shifted maximum. This difference is not due to changes

in the lowest energy pigments, but rather in the environment of Chls absorbing around 674

nm, which in plants are located in the 602 and 603 binding sites (Remelli, Varotto et al.

1999).

4.4. Energy transfer in PSII-LHCII of C. reinhardtii

The role of LHCII is to absorb light and to transfer excitation energy to the reaction center

where charge separation occurs. The connection of LHCII trimers with the core is thus

essential in assuring efficient energy transfer (Croce and van Amerongen 2011). In higher

plants it has been shown that the overall trapping time in the PSII supercomplex is 143 ps

(Caffarri, Broess et al. 2011) and that this very fast transfer is probably due to the presence

of preferential energy transfer pathways between complexes involving mainly Chls a.

The reconstituted structure of the C.r.PSII supercomplex gives the possibility to point out

preferential energy transfer pathways, based on the organization of the Chls. The model of

Page 90: University of Groningen From Photosystem I to Photosystem

89

the supercomplex shows that the relative orientation of the S trimer is identical to that in

higher plants, which means that chlorophylls 610-611-612 (nomenclature from (Liu, Yan et

al. 2004)) of one LHCII monomer, are close to Chl 633 of the core, (nomenclature from

(Umena, Kawakami et al. 2011)), allowing for rapid energy transfer as in plants.

The orientation of trimer M is different than in A. thaliana. In this configuration Chls a 610

and 604 of LHCII face directly CP29, while Chls 611 and 612 of a different monomer face

trimer N. In the case of trimer N the shorter Chl (LHCII)-Chl(core) distance (around 20 Å) is

between Chls 611 and 612 of one LHCII trimer and Chl 612 of the core. In conclusion, also in

the case of C. reinhardtii Chls 610-612 seem to be involved in the transfer between

complexes, although the connection between trimer M and CP29 involves Chl 610 instead of

Chls 611 and 612 as in A. thaliana. These Chls represent the lowest energy sites in LHCII of

higher plants (Remelli, Varotto et al. 1999; Novoderezhkin, Palacios et al. 2005) and they are

likely to be conserved also in C. reinhardtii. The minimal distance between Chls in trimer N

and in the core is rather large, suggesting a slower transfer, although at this stage we cannot

exclude that an additional Chl could be located in between these two complexes in C.r.

4.5. The thylakoid membrane of C. reinhardtii harbors at least 6 LHCII trimers per

monomeric PSII core complex

The fraction containing the largest PSII supercomplex has a Chl a/b ratio of 2.22, in perfect

agreement with the presence of 3.7 trimers on average per dimeric core complex in this

fraction. This number is very close to the Chl a/b ratio of C.reinhardtii cells which is 2.28.

However, the cells also contain PSI in a close to 1:1 stoichiometry with PSII, in agreement

with previous studies (Melis, Murakami et al. 1996; Polle, Benemann et al. 2000; Petroutsos,

Terauchi et al. 2009; Yan, Schofield et al. 2011). Considering that PSI coordinates around

240 Chls, 196 Chl a and 44 Chl b (Drop, Webber-Birungi et al. 2011), the low Chl a/b ratio of

the cells can only be explained with the presence of extra LHCII trimers in the

photosynthetic membrane of C.reinhardtii. The number can be roughly calculated on the

basis of the pigment stoichiometry of core (35 Chls a) and PSII antennas (Hogewoning,

Wientjes et al. 2012). We have used 14 Chls a value of 1.28 for the Chl a/b (than 7.85 Chl a

and 6.15 Chl b) of all antenna complexes, although it is probably on the low side for CP29

and CP26, which means that we underestimate the number of LHCII trimers. The number X

of antenna (monomers) is thus calculated as 2.28 = (196 + 35 + 7.85X)/(44 + 6.15X).

Page 91: University of Groningen From Photosystem I to Photosystem

90

The result suggests the presence of 3-4 additional LHCII trimers per monomeric core

(meaning 6/7 in total), making the light harvesting capacity of C.reinhardtii far larger than

that of plants, in which a maximum of 4.5 LHCII trimers was observed (Peter and Thornber

1991; van Oort, Alberts et al. 2010). Where are the additional trimers located? Looking at

the structure of the PSII supercomplex it is clear that there is space at most for one extra

trimer associated with the core. The tighter possible organization of the PSII

supercomplexes of C.reinhardtii is shown in Fig. 8 and indicates that differently from the

case of plants the dimeric cores cannot be directly connected but are separated by

a minimum of one row of LHCII. In this case we can speculate that trimer N should be able

to transfer energy to two different core complexes.

It remains to be elucidated if the “extra” trimers are located in between the PSII

supercomplexes and if a subpopulation of them is associated with PSI in all conditions, as it

is the case in plants (Wientjes, van Amerongen et al. 2013) or if they form LHCII-only

domains.

Figure 8. Model of the tighter possible organization of the C2S2M2N2 supercomplexes of Chlamydomonas reinhardtii.

Acknowledgments

We thank Kevin Redding (Arizona State University, Tempe, USA) for kindly providing

the C. reinhardtii strain (JVD-1B [pGG1]) used in this work and Pierre Cardol (Université de

Liège) for help with the ECS measurements. This work was supported by the ERC Starting

Grant 281341 (ASAP) to RC and FOM grant 10TBSC12-2 to EJB.

Page 92: University of Groningen From Photosystem I to Photosystem

91

References

Angeler, D. G. and M. Schagerl (1997). "Distribution of the xanthophyll loroxanthin in selected members of the Chlamydomonadales and Volvocales (Chlorophyta)." Phyton-Annales Rei Botanicae 37(1): 119-132.

Bailleul, B., P. Cardol, et al. (2010). "Electrochromism: a useful probe to study algal photosynthesis." Photosynthesis Research 106(1-2): 179-189.

Bassi, R. (1985). "Spectral Properties and Polypeptide Composition of the Chlorophyll-Proteins from Thylakoids of Granal and Agranal Chloroplasts of Maize (Zea-Mays-L)." Carlsberg Research Communications 50(2): 127-143.

Bassi, R., B. Pineau, et al. (1993). "Carotenoid-binding proteins of photosystem II." European Journal of Biochemistry 212(2): 297-303.

Bassi, R. and F. A. Wollman (1991). "The Chlorophyll-a/B Proteins of Photosystem-Ii in Chlamydomonas-Reinhardtii - Isolation, Characterization and Immunological Cross-Reactivity to Higher-Plant Polypeptides." Planta 183(3): 423-433.

Ben-Shem, A., F. Frolow, et al. (2003). "Crystal structure of plant photosystem I." Nature 426(6967): 630-635.

Berthold, D. A., G. T. Babcock, et al. (1981). "A Highly Resolved, Oxygen-Evolving Photosystem-Ii Preparation from Spinach Thylakoid Membranes - Electron-Paramagnetic-Res and Electron-Transport Properties." Febs Letters 134(2): 231-234.

Betterle, N., M. Ballottari, et al. (2009). "Light-induced Dissociation of an Antenna Hetero-oligomer Is Needed for Non-photochemical Quenching Induction." Journal of Biological Chemistry 284(22): 15255-15266.

Boekema, E. J., H. van Roon, et al. (1999). "Supramolecular organization of photosystem II and its light-harvesting antenna in partially solubilized photosystem II membranes." European Journal of Biochemistry 266(2): 444-452.

Buchel, C. and W. Kuhlbrandt (2005). "Structural differences in the inner part of Photosystem II between higher plants and cyanobacteria." Photosynthesis Research 85(1): 3-13.

Caffarri, S., K. Broess, et al. (2011). "Excitation Energy Transfer and Trapping in Higher Plant Photosystem II Complexes with Different Antenna Sizes." Biophysical Journal 100(9): 2094-2103.

Caffarri, S., R. Kouril, et al. (2009). "Functional architecture of higher plant photosystem II supercomplexes." Embo Journal 28(19): 3052-3063.

Caffarri, S., F. Passarini, et al. (2007). "A specific binding site for neoxanthin in the monomeric antenna proteins CP26 and CP29 of Photosystem II." Febs Letters 581(24): 4704-4710.

Croce, R., G. Canino, et al. (2002). "Chromophore organization in the higher-plant photosystem II antenna protein CP26." Biochemistry 41(23): 7334-7343.

Croce, R. and H. van Amerongen (2011). "Light-harvesting and structural organization of Photosystem II: From individual complexes to thylakoid membrane." Journal of Photochemistry and Photobiology B-Biology 104(1-2): 142-153.

Dainese, P. and R. Bassi (1991). "Subunit Stoichiometry of the Chloroplast Photosystem-Ii Antenna System and Aggregation State of the Component Chlorophyll-a/B Binding-Proteins." Journal of Biological Chemistry 266(13): 8136-8142.

Dang, N. C., V. Zazubovich, et al. (2008). "The CP43 proximal antenna complex of higher plant photosystem II revisited: Modeling and hole burning study. I." Journal of Physical Chemistry B 112(32): 9921-9933.

Dau, H., I. Zaharieva, et al. (2012). "Recent developments in research on water oxidation by photosystem II." Curr Opin Chem Biol 16(1-2): 3-10.

de Bianchi, S., M. Ballottari, et al. (2010). "Regulation of plant light harvesting by thermal dissipation of excess energy." Biochem Soc Trans 38(2): 651-660.

Page 93: University of Groningen From Photosystem I to Photosystem

92

de Bianchi, S., L. Dall'Osto, et al. (2008). "Minor antenna proteins CP24 and CP26 affect the interactions between photosystem II Subunits and the electron transport rate in grana membranes of Arabidopsis." Plant Cell 20(4): 1012-1028.

Dekker, J. P. and E. J. Boekema (2005). "Supramolecular organization of thylakoid membrane proteins in green plants." Biochimica Et Biophysica Acta-Bioenergetics 1706(1-2): 12-39.

Drop, B., M. Webber-Birungi, et al. (2011). "Photosystem I of Chlamydomonas reinhardtii Contains Nine Light-harvesting Complexes (Lhca) Located on One Side of the Core." Journal of Biological Chemistry 286(52): 44878-44887.

Elrad, D. and A. R. Grossman (2004). "A genome's-eye view of the light-harvesting polypeptides of Chlamydomonas reinhardtii." Current Genetics 45(2): 61-75.

Elrad, D., K. K. Niyogi, et al. (2002). "A major light-harvesting polypeptide of photosystem II functions in thermal dissipation." Plant Cell 14(8): 1801-1816.

Ferrante, P., M. Ballottari, et al. (2012). "LHCBM1 and LHCBM2/7 polypeptides, components of major LHCII complex, have distinct functional roles in photosynthetic antenna system of Chlamydomonas reinhardtii." Journal of Biological Chemistry 287(20): 16276-16288.

Fischer, N., P. Setif, et al. (1997). "Targeted mutations in the psaC gene of Chlamydomonas reinhardtii: Preferential reduction of F-B at low temperature is not accompanied by altered electron flow from photosystem I ferredoxin." Biochemistry 36(1): 93-102.

Gorman, D. S. and R. P. Levine (1965). "Cytochrome f and plastocyanin: their sequence in the photosynthetic electron transport chain of Chlamydomonas reinhardi." Proc Natl Acad Sci U S A 54(6): 1665-1669.

Goussias, C., A. Boussac, et al. (2002). "Photosystem II and photosynthetic oxidation of water: an overview." Philosophical Transactions of the Royal Society B-Biological Sciences 357(1426): 1369-1381.

Grossman, A. R., M. Lohr, et al. (2004). "Chlamydomonas reinhardtii in the landscape of pigments." Annual Review of Genetics 38: 119-173.

Gulis, G., K. V. Narasimhulu, et al. (2008). "Purification of His(6)-tagged photosystem i from Chlamydomonas reinhardtii." Photosynthesis Research 96(1): 51-60.

Guskov, A., J. Kern, et al. (2009). "Cyanobacterial photosystem II at 2.9-angstrom resolution and the role of quinones, lipids, channels and chloride." Nat Struct Mol Biol 16(3): 334-342.

Hogewoning, S. W., E. Wientjes, et al. (2012). "Photosynthetic Quantum Yield Dynamics: From Photosystems to Leaves." Plant Cell 24(5): 1921-1935.

Ifuku, K., S. Ishihara, et al. (2010). "Molecular functions of oxygen-evolving complex family proteins in photosynthetic electron flow." J Integr Plant Biol 52(8): 723-734.

Iwai, M., Y. Takahashi, et al. (2008). "Molecular remodeling of photosystem II during state transitions in Chlamydomonas reinhardtii." Plant Cell 20(8): 2177-2189.

Jansson, S. (1999). "A guide to the Lhc genes and their relatives in Arabidopsis." Trends in Plant Science 4(6): 236-240.

Joliot, P. and R. Delosme (1974). "Flash-Induced 519nm Absorption Change in Green-Algae." Biochimica Et Biophysica Acta 357(2): 267-284.

Knoetzel, J., T. Braumann, et al. (1988). "Pigment Protein Complexes of Green-Algae - Improved Methodological Steps for the Quantification of Pigments in Pigment Protein Complexes Derived from the Green-Algae Chlorella and Chlamydomonas." Journal of Photochemistry and Photobiology B-Biology 1(4): 475-491.

Kovacs, L., J. Damkjaer, et al. (2006). "Lack of the light-harvesting complex CP24 affects the structure and function of the grana membranes of higher plant chloroplasts." Plant Cell 18(11): 3106-3120.

Liu, H. B., R. G. Sadygov, et al. (2004). "A model for random sampling and estimation of relative protein abundance in shotgun proteomics." Analytical Chemistry 76(14): 4193-4201.

Liu, Z. F., H. C. Yan, et al. (2004). "Crystal structure of spinach major light-harvesting complex at 2.72 angstrom resolution." Nature 428(6980): 287-292.

Page 94: University of Groningen From Photosystem I to Photosystem

93

Melis, A., A. Murakami, et al. (1996). "Chromatic regulation in Chlamydomonas reinhardtii alters photosystem stoichiometry and improves the quantum efficiency of photosynthesis." Photosynthesis Research 47(3): 253-265.

Merchant, S. S., S. E. Prochnik, et al. (2007). "The Chlamydomonas genome reveals the evolution of key animal and plant functions." Science 318(5848): 245-251.

Minagawa, J. and Y. Takahashi (2004). "Structure, function and assembly of Photosystem II and its light-harvesting proteins." Photosynthesis Research 82(3): 241-263.

Muller, P., X. P. Li, et al. (2001). "Non-photochemical quenching. A response to excess light energy." Plant Physiology 125(4): 1558-1566.

Nield, J., O. Kruse, et al. (2000). "Three-dimensional structure of Chlamydomonas reinhardtii and Synechococcus elongatus photosystem II complexes allows for comparison of their oxygen-evolving complex organization." Journal of Biological Chemistry 275(36): 27940-27946.

Nield, J., E. V. Orlova, et al. (2000). "3D map of the plant photosystem II supercomplex obtained by cryoelectron microscopy and single particle analysis." Nat Struct Biol 7(1): 44-47.

Novoderezhkin, V. I., M. A. Palacios, et al. (2005). "Excitation dynamics in the LHCII complex of higher plants: Modeling based on the 2.72 angstrom crystal structure." Journal of Physical Chemistry B 109(20): 10493-10504.

Oostergetel, G. T., W. Keegstra, et al. (1998). "Automation of specimen selection and data acquisition for protein electron crystallography." Ultramicroscopy 74(1-2): 47-59.

Pan, X. W., M. Li, et al. (2011). "Structural insights into energy regulation of light-harvesting complex CP29 from spinach." Nat Struct Mol Biol 18(3): 309-U394.

Passarini, F., E. Wientjes, et al. (2009). "Molecular basis of light harvesting and photoprotection in CP24: unique features of the most recent antenna complex." Journal of Biological Chemistry 284(43): 29536-29546.

Peter, G. F. and J. P. Thornber (1991). "Biochemical-Composition and Organization of Higher-Plant Photosystem-Ii Light-Harvesting Pigment-Proteins." Journal of Biological Chemistry 266(25): 16745-16754.

Petroutsos, D., A. M. Terauchi, et al. (2009). "PGRL1 Participates in Iron-induced Remodeling of the Photosynthetic Apparatus and in Energy Metabolism in Chlamydomonas reinhardtii." Journal of Biological Chemistry 284(47): 32770-32781.

Pineau, B., C. Gerard-Hirne, et al. (2001). "Carotenoid binding to photosystems I and II of Chlamydomonas reinhardtii cells grown under weak light or exposed to intense light." Plant Physiology and Biochemistry 39(1): 73-85.

Polle, J. E. W., J. R. Benemann, et al. (2000). "Photosynthetic apparatus organization and function in the wild type and a chlorophyll b-less mutant of Chlamydomonas reinhardtii. Dependence on carbon source." Planta 211(3): 335-344.

Porra, R. J., W. A. Thompson, et al. (1989). "Determination of Accurate Extinction Coefficients and Simultaneous-Equations for Assaying Chlorophyll-a and Chlorophyll-B Extracted with 4 Different Solvents - Verification of the Concentration of Chlorophyll Standards by Atomic-Absorption Spectroscopy." Biochimica Et Biophysica Acta 975(3): 384-394.

Remelli, R., C. Varotto, et al. (1999). "Chlorophyll binding to monomeric light-harvesting complex - A mutation analysis of chromophore-binding residues." Journal of Biological Chemistry 274(47): 33510-33521.

Shi, L. X., M. Hall, et al. (2012). "Photosystem II, a growing complex: Updates on newly discovered components and low molecular mass proteins." Biochimica Et Biophysica Acta-Bioenergetics 1817(1): 13-25.

Standfuss, R., A. C. T. van Scheltinga, et al. (2005). "Mechanisms of photoprotection and nonphotochemical quenching in pea light-harvesting complex at 2.5A resolution." Embo Journal 24(5): 919-928.

Stauber, E. J., A. Fink, et al. (2003). "Proteomics of Chlamydomonas reinhardtii light-harvesting proteins." Eukaryotic Cell 2(5): 978-994.

Page 95: University of Groningen From Photosystem I to Photosystem

94

Takahashi, H., M. Iwai, et al. (2006). "Identification of the mobile light-harvesting complex II polypeptides for state transitions in Chlamydomonas reinhardtii." Proceedings of the National Academy of Sciences of the United States of America 103(2): 477-482.

Teramoto, H., T. Ono, et al. (2001). "Identification of Lhcb gene family encoding the light-harvesting chlorophyll-a/b proteins of photosystem II in Chlamydomonas reinhardtii." Plant and Cell Physiology 42(8): 849-856.

Tokutsu, R., N. Kato, et al. (2012). "Revisiting the Supramolecular Organization of Photosystem II in Chlamydomonas reinhardtii." Journal of Biological Chemistry 287(37): 31574-31581.

Umena, Y., K. Kawakami, et al. (2011). "Crystal structure of oxygen-evolving photosystem II at a resolution of 1.9 angstrom." Nature 473(7345): 55-U65.

van Oort, B., M. Alberts, et al. (2010). "Effect of Antenna-Depletion in Photosystern II on Excitation Energy Transfer in Arabidopsis thaliana." Biophysical Journal 98(5): 922-931.

Wientjes, E., H. van Amerongen, et al. (2013). "LHCII is an antenna of both photosystems after long-term acclimation." Biochimica Et Biophysica Acta.

Yakushevska, A. E., P. E. Jensen, et al. (2001). "Supermolecular organization of photosystem II and its associated light-harvesting antenna in Arabidopsis thaliana." European Journal of Biochemistry 268(23): 6020-6028.

Yakushevska, A. E., W. Keegstra, et al. (2003). "The structure of photosystem II in Arabidopsis: Localization of the CP26 and CP29 antenna complexes." Biochemistry 42(3): 608-613.

Yan, C. Y., O. Schofield, et al. (2011). "Photosynthetic energy storage efficiency in Chlamydomonas reinhardtii, based on microsecond photoacoustics." Photosynthesis Research 108(2-3): 215-224.

Page 96: University of Groningen From Photosystem I to Photosystem

95

CHAPTER 4

Consequences of state transitions on the structural and functional

organization of Photosystem I in the green alga Chlamydomonas

reinhardtii

Bartlomiej Drop1, Sathish Yadav K.N.2, Egbert J. Boekema2 and Roberta Croce1

1 Department of Physics and Astronomy, Faculty of Sciences, VU University Amsterdam,

De Boelelaan 1081, 1081 HV Amsterdam, The Netherlands.

2 Department of Electron Microscopy, Groningen Biological Sciences and Biotechnology

Institute, University of Groningen, Nijenborgh 7, 9747 AG Groningen, The Netherlands.

Accepted for pubblication in The Plant Journal

Page 97: University of Groningen From Photosystem I to Photosystem

96

Abstract

State transitions represent a photoacclimation process that regulates the light-driven

photosynthetic reactions in response to changes in light quality/quantity. It balances the

excitation between Photosystem I (PSI) and II (PSII) by shuttling LHCII, the main light-

harvesting complex of green algae and plants, between them. This process is particularly

important in Chlamydomonas renihardtii where it is suggested to induce a large

reorganization in the thylakoid membrane. Phosphorylation has been shown to be

necessary for state transitions and the LHCII kinase has been identified. However, the

consequences of state transitions on the structural organization and the functionality of the

photosystems have not yet been elucidated. This is mainly because the purification of the

supercomplexes has proved to be particularly difficult, thus preventing structural and

functional studies. Here, we have purified and analysed PSI and PSII supercomplexes of C.

reinhardtii in state1 and 2, and have studied them using biochemical, spectroscopic and

structural methods. It is shown that PSI in state2 is able to bind two LHCII trimers containing

all four LHCII types, and one monomer, most likely CP29, in addition to its nine Lhcas. This is

the largest PSI complex ever observed, having an antenna size of 340 Chls/P700. Moreover,

all PSI-bound Lhcs are efficient in transferring energy to PSI. A projection map at 20 Å

resolution reveals the structural organization of the complex. Surprisingly, only LHCII type I,

II and IV are phosphorylated when associated with PSI, while LHCII type II and CP29 are not,

but CP29 is phosphorylated when associated with PSII in state2.

Page 98: University of Groningen From Photosystem I to Photosystem

97

1. Introduction

Photosynthetic light harvesting is carried out by two pigment-binding multiprotein

complexes, Photosystem II (PSII) and Photosystem I (PSI), which work in series to drive

electrons from water to NADPH. To maintain optimal linear electron transport and to avoid

photodamage, it is important that an excitation balance is maintained between the two

photosystems. However, PSI and PSII have slightly different characteristics, with the former

absorbing far-red light more efficiently (around 690 nm) and the latter more red light

(around 650 nm) (Rochaix 2007). Moreover, they also differ in their quantum efficiency of

charge separation which is close to 1 for PSI (Nelson 2009) and rather variable for PSII,

depending on the antenna size (Wientjes, van Amerongen et al. 2013). In plants and green

algae, PSI and PSII are composed of two moieties: the core complex that contains all the

cofactors of the election transport chain, and the outer antenna system, which extends the

light-harvesting capacity of the core and ensures photoprotection (Dekker and Boekema

2005; Croce and van Amerongen 2011). The outer antenna is composed of members of the

Lhc (Light-harvesting complex) gene family, which are structurally highly homologous: each

Lhc apoprotein has three transmembrane α-helices and coordinates Chlorophylls (Chl) a,

Chls b, and different carotenoid molecules (Ben-Shem, Frolow et al. 2003; Liu, Yan et al.

2004; Standfuss, van Scheltinga et al. 2005). In plants, six Lhca complexes and six Lhcb

complexes which serve as antenna of PSI and PSII respectively (Jansson 1999). In the green

alga model organism C. reinhardtii, there are twenty Lhc antennas genes encoding nine

Lhcas and eleven Lhcbs, nine of which (LhcbM1-9) codify for the major antenna complex

LHCII and two for the monomeric antennas CP26 (Lhcb5) and CP29 (Lhcb4) (Elrad and

Grossman 2004; Merchant, Prochnik et al. 2007). The antenna size of both PSI and PSII was

shown to be larger in C. reinhardtii as compared to higher plants (Germano, Yakushevska et

al. 2002; Kargul, Nield et al. 2003; Stauber, Busch et al. 2009; Drop, Webber-Birungi et al.

2011; Drop, Webber-Birungi et al. 2014).

In order to survive and grow optimally, photosynthetic organisms need to acclimate to

constantly changing light conditions (Rochaix 2007). One of the strategies for light

acclimation at the level of the photosynthetic apparatus relies on structural rearrangements

of the peripheral antenna system which occur on different time scales (Bellafiore, Barneche

et al. 2005; Tikkanen, Pippo et al. 2006). Short term acclimation to changes in illumination

triggers state transitions. In this process, energy distribution between photosystems is

Page 99: University of Groningen From Photosystem I to Photosystem

98

balanced by the reversible association of LHCII with either PSII (state 1) or PSI (state 2)

(Bonaventura and Myers 1969; Murata 1969; Allen and Forsberg 2001). In PSII-favoring light

conditions, the plastoquinone (PQ) pool is reduced (Allen, Bennett et al. 1981), and

plastoquinol docks to the Qo site of the cytochrome b6f complex (Zito, Finazzi et al. 1999)

triggering the activation of a thylakoid-bound kinase (Stt7 C. reinhardtii and its orthologue

STN7 in higher plant) (Bellafiore, Barneche et al. 2005; Bonardi, Pesaresi et al. 2005;

Lemeille, Willig et al. 2009) that phosphorylates LHCII (Bennett, Steinback et al. 1980;

Horton and Black 1980; Depege, Bellafiore et al. 2003; Bellafiore, Barneche et al. 2005). As a

consequence, LHCII detaches from PSII and migrates to PSI (Andersson, Akerlund et al. 1982;

Kyle, Staehelin et al. 1983). Reversely, in PSI-favoring light conditions, the PQ pool is

oxidized, leading to dephosphorylation of LHCII (by TAP38/PPH1 phosphatase in plants), its

dissociation from PSI and reassociation with PSII (Bennett 1980; Pribil, Pesaresi et al. 2010;

Shapiguzov, Ingelsson et al. 2010). In most light conditions, plant LHCII is associated with

both PSI and PSII in an amount that depends on the light intensity, thus allowing also for

long term acclimation by regulating only the expression of Lhcb1 and Lhcb2 genes (Wientjes,

van Amerongen et al. 2013). In higher plants only a limited fraction of LHCII is involved in

state transitions (Allen 1992), whereas in C.reinhardtii it was reported that 80% of the LHCII

pool (Delosme, Olive et al. 1996) and the monomeric complexes CP29 and CP26 (Kargul,

Turkina et al. 2005; Takahashi, Iwai et al. 2006; Tokutsu, Iwai et al. 2009), participate in this

process and reversibly shuttle between PSII and PSI, inducing a large membrane

reorganization (Chuartzman, Nevo et al. 2008; Minagawa 2011). It has been proposed that

photosystem II supercomplexes in particular undergo strong remodeling during transitions

from state 1 to state 2 (Iwai, Takahashi et al. 2008).

In higher plants, it has been shown that one LHCII trimer is associated with the PSI-LHCI

complex on the side of PsaH and PsaL (Lunde, Jensen et al. 2000; Kouril, Zygadlo et al. 2005)

to which it transfers excitation energy very efficiently (Galka, Santabarbara et al. 2012;

Wientjes, van Amerongen et al. 2013). In state 1 the same LHCII acts as an antenna of PSII

but shows lower transfer efficiency (Wientjes, van Amerongen et al. 2013). Although the

extent of state transitions is considered to be very large in C.reinhardtii (Delosme, Olive et

al. 1996), the purification of the PSI-LHCII particles has proven to be very difficult. A particle

containing CP29 with only four of the nine Lhcas has been purified and analysed by EM by

Kargul et al. (Kargul, Turkina et al. 2005), while Takahashi et al. (Takahashi, Iwai et al. 2006)

Page 100: University of Groningen From Photosystem I to Photosystem

99

reported the purification of a PSI particle containing CP29, CP26 and LhcbM type II, but no

information regarding the stoichiometry of the subunits, the size, or the functional and

structural organization of the supercomplex is currently available.

In this study, using a milder solubilization procedure than any previously reported, we have

purified the PSI-LHCI-LHCII supercomplex from C. reinhardtii and characterized it by

combining biochemistry, spectroscopy and electron microscopy measurements. The

complex is far larger than expected, and the EM projection map clearly shows in addition to

all nine Lhca complexes, two LHCII trimers, and one monomeric Lhcb complex which

connects one of the trimers to the PSI core. Likewise, the protein composition differs from

what was previously observed; in addition to CP29, CP26 and LhcbM type II (LhcbM5) as

reported by Takahashi et al. (Takahashi, Iwai et al. 2006), LhcbM type I (LhcbM3/4/6/8/9), III

(LhcbM2/7) and IV (LhcbM1) are also associated with PSI. We furthermore prove that the

association of Lhcb complexes to PSI is functional. These results, together with the analysis

of PSII-LHCII complexes purified in state 2, allow a series of open questions regarding the

functional and structural organization of the photosystems during state transitions in

C. reinhardtii to be addressed.

2. Materials and methods

2.1. Strain and growth conditions

Chlamydomonas reinhardtii strains JVD-1B[pGG1] (kind gift of K. Redding), in which a hexa-

histidine tag has been added at the N-terminus of PsaA core subunit of PSI (Gulis,

Narasimhulu et al. 2008), and the PSII mutant (Fl39) (Maroc, Garnier et al. 1989) were

grown in liquid Tris-Acetate-Phosphate (TAP) medium (Gorman and Levine 1965). Cells were

cultured at room temperature (25°C) on an incubator shaker (Minitron, INFORS HT) at 170

rpm under a continuous illumination flux of 20 μmol photons PAR m-2 s-1.

2.2 Induction of state transitions and thylakoid membranes preparation

C.reinhardtii cells were grown until the mid-logarithmic phase (OD750nm ≈ 0.7–0.8). To induce

state 2 C.reinhardtii cells were incubated in anaerobic conditions in the dark for 20 min in

the presence of 0.5 mM NaN3, (Rebeille and Gans 1988). To induce state 1 cells were

incubate with 10 μM 3,4-dichlorophenyl-1, 1-dimethylurea (DCMU) for 20 min in light.

Page 101: University of Groningen From Photosystem I to Photosystem

100

2.3 Thylakoid preparation

The C.reinhardtii cultures were harvested (3000 rpm, 5 min, 4°C) and resuspended in 25 mM

Mes, 0.33 M sucrose, 5 mM MgCl2, 1.5 mM NaCl, 1 mM aminocaproic acid,

1 mM aminobenzamidine (pH 6.5) buffer. Cells were disrupted by sonication (60W power in

10 cycles of 10 s on/30 s off) and diluted 5 times with 25mM Hepes (pH 7.5), 0.33 M

sucrose, 5 mM EDTA buffer. Membranes were harvested by centrifugation at 17000 rpm for

20 min and unbroken cells were removed from the membrane fraction by centrifugation at

3000 rpm for 30 s. Thylakoids were harvested following centrifugation at 11,000 rpm for

12 min and resuspended in 25mM Hepes (pH 7.5), 50% glycerol (w/v), 5mM MgCl2 buffer.

The buffers used for preparation of thylakoids membranes in state 2 contained 0.5 mM

NaN3.

2.4 Isolation of PSI-LHCI-LHCII supercomplexes

PSI-LHCI-LHCII isolation was modified from (Drop, Webber-Birungi et al. 2011). Thylakoids

were pelleted, unstacked with 5 mM EDTA and washed with 20 mM Hepes (pH 7.5).

Membranes were then resuspended in 20 mM Hepes (pH 7.5), 0.15 M NaCl and solubilized

at a final chlorophyll concentration of 0.5 mg/ml by adding an equal volume of 1% digitonin

and 0.2% α-dodecylmaltoside (α-DM) (Galka, Santabarbara et al. 2012). Samples were

vortexed for a few seconds and then centrifuged (12000 rpm, 10 min, 4 °C) to remove

unsolubilized material. The buffers used for isolation of the PSI supercomplex in state 2

contained 0.5 mM NaN3.

In the case of thylakoids from the JVD-1B[pGG1] cells, the supernatant was loaded onto

a HisTrap HP Column (GE Healthcare) equilibrated with 20 mM Hepes (pH 7.5), 0.15 M NaCl,

0.02% digitonin and washed with the same buffer. Elution was carried out with 20 mM

Hepes (pH 7.5), 0.2 M imidazole, 0.15 M NaCl, 0.02% digitonin. The eluted fraction was

loaded on a sucrose density gradient (made by freezing and thawing 0.6 M sucrose, 20 mM

Hepes (pH 7.5), 0.02% digitonin buffer layered over 1ml of 2M sucrose) and purified to

homogeneity by ultracentrifugation (41000 rpm, 12 h, 4 °C). PSI supercomplexes were

collected using a syringe and directly used for measurements or rapidly frozen in liquid

nitrogen.

The purification of PSI supercomplexes from Fl39 (PSII mutant) was performed under the

same conditions as used for the purification of the PSI supercomplexes described above,

Page 102: University of Groningen From Photosystem I to Photosystem

101

with the difference that upon solubilization the membranes were directly loaded on

a sucrose gradient.

2.5 Isolation of the PSII-LHCII supercomplexes

For the purification of PSII supercomplexes in state 1 and 2 the flow through of the HisTrap

column was loaded on a sucrose density gradient (prepared by freezing and thawing 0.5 M

sucrose, 20 mM Hepes (pH 7.5), 0.02% digitonin layered over 1ml of 2M sucrose). PSII

complexes were separated by ultracentrifugation (41000 rpm, 14 h, 4°C). The green bands

visible on the sucrose gradient were harvested with a syringe.

2.6. Gel electrophoresis

Proteins were analyzed by SDS-6M urea PAGE with a Tris-sulphate buffer system (Bassi

1985) at 14% acrylamide concentration. The amounts of sample loaded into each well were:

3 µg (in Chls) for thylakoids; 2.5 µg for PSI and PSII supercomplexes; 1.5 µg for LHCII and

Lhcb fractions. Gels were stained with 0.025% Coomassie Blue G in 50% methanol and 10%

acetic acid. Pro-Q Diamond Phosphoprotein Gel Stain (Molecular Probes) was used as

described in the user manual, except that the samples were not desalted and delipidated.

Immunoblot analysis was performed against CP26 and CP29 using antibodies from Agrisera.

The membranes were exposed to SuperSignal West Pico Chemiluminescent Substrate

(Thermo Scientific).

The Coomassie and Pro-Q Diamond stained gels and blot membranes were imaged with

ImageQuant LAS-4000 (GE Healthcare).

2.7 Pigment composition

The chlorophyll concentrations of thylakoid preparations were calculated in 80% (v/v)

acetone, according to (Porra, Thompson et al. 1989). The pigment composition of the

complexes was analyzed by fitting the spectrum of the 80% acetone extracted pigments

with the spectra of the individual pigments in acetone and by HPLC, as described previously

(Croce, Canino et al. 2002). The data are the results of at least three different preparations

in two replicas.

2.8 Spectroscopic analysis

Room-temperature and 77K absorption spectra were recorded with a Cary 4000

spectrophotometer (Varian). The fluorescence emission spectra were recorded at 5°C and

Page 103: University of Groningen From Photosystem I to Photosystem

102

77K using a Fluorolog 3.22 spectrofluorimeter (Jobin Yvon-Spex). The 77K measurements

were carried out in a home-built liquid nitrogen cooled device or in an Oxford liquid

nitrogen bath cryostat (77 K). Before measurements cells were diluted in TAP medium

containing 66.7% glycerol (w/v). All fluorescence spectra were measured at OD 0.05 at the

maximum of the Qy absorption.

Room temperature absorption measurements were performed in 0.5 M sucrose, 20 mM

Hepes (pH 7.5), 0.02% digitonin buffer.

2.9 Electron microscopy and single particle analysis

Samples from the sucrose density centrifugation were negatively stained using the droplet

method with 2% uranyl acetate on glow discharged carbon-coated copper grids. Electron

microscopy was performed on a Philips CM120 electron microscope equipped with a LaB6

filament operating at 120 kV. Images were recorded with a Gatan 4000 SP 4K slow-scan CCD

camera at either 130000 × magnification at a pixel size (after binning the images) of 2.25 Å,

respectively, at the specimen level with GRACE software for semi-automated specimen

selection and data acquisition (Oostergetel, Keegstra et al. 1998). Single particle analysis

was performed using GRIP software including multi-reference and non-reference

alignments, multivariate statistical analysis, and classification, as in (Boekema, van Roon et

al. 1999).

3. Results

3.1 PSI-LHCI-LHCII supercomplex can be purified from state 2 cells

The PSI-LHCI-LHCII supercomplex was purified from two different modified strains in order

to avoid PSII contamination: the WT JDS-1B-His (in the following called WT-His), carrying

a histidine tag at the N-terminus of the core subunit PsaA (Gulis, Narasimhulu et al. 2008),

and the Fl39 mutant, which lacks the PSII core, but still contains Lhcb proteins and is able to

respond to state 1 and state 2 conditions as shown by low temperature fluorescence

(Maroc, Garnier et al. 1989). The comparison of the complexes obtained from the two

strains also allows us to check for a possible effect of the strain itself on the complex. Cells

were locked in state 2 by incubation in anaerobic conditions in the dark, to reduce the

plastoquinone pool (Bulte, Gans et al. 1990), and in the presence of sodium azide, which

inhibits mitochondrial respiration. This method is commonly used to induce state transition

Page 104: University of Groningen From Photosystem I to Photosystem

103

in this alga (see for e.g (Zito, Finazzi et al. 1999; Depege, Bellafiore et al. 2003; Lemeille,

Willig et al. 2009; Takahashi, Clowez et al. 2013).

Figure 1. Fluorescence emission spectra of the cells. Fluorescence emission spectra of the cells of WT-His (A) and Fl39 (B) were measured at 77 K without any treatment (black) or after induction of state 1 (light grey) or state 2 (dark grey). The excitation wavelength was 475 nm. The spectra are normalized to the maximum.

Cells grown in the presence of light and oxygen were used as a control. In control

conditions, the WT-His cells were in state 1 as indicated by the fact that their fluorescence

emission spectrum at 77K was identical to that of the cells locked in state 1 by addition of

DCMU (Fig. 1A). The fluorescence emission spectrum of the Fl39 mutant in control growth

conditions showed a small difference with respect to the Fl39 cells in which state 1 was

induced (Fig. 1B), suggesting that part of the LHCII was associated with PSI in this mutant in

growth conditions.

Fluorescence emission spectra at 77K of the thylakoid membranes (Fig. 2) showed that the

membrane were still in state 1 and 2 upon purification. The thylakoid membranes isolated

from WT-His and Fl39 cells were mildly solubilized with digitonin and loaded on a sucrose

gradient either directly (Fl39 thylakoids) or after elution from a Ni-Sepharose column (WT-

His thylakoids, see experimental procedures for details). The affinity chromatography step

used for the WT-His thylakoids takes advantage of the His-tag on PsaA to separate PSI from

Page 105: University of Groningen From Photosystem I to Photosystem

104

PSII sub- and supercomplexes (Gulis et al. 2008; Drop et al. 2011). The sucrose gradients are

shown in Fig. 3A. In the gradient loaded with WT-His PSI complexes, two bands were visible

in state 1 and three in state 2. In addition, in state 2 a band was observed at the bottom of

the gradient (Fig. 3A), likely representing partially insolubilized material due to the very mild

solubilization conditions. In the gradient loaded with the solubilized thylakoids isolated from

the Fl39 strain, in addition to the bands containing Lhcb complexes (B1-B3), two bands (B4

and B5) were also visible in the lower part of the gradients in both state 1 and 2 (Fig. 3A).

The absorption spectra of the bands were measured and a selection is shown in Fig. 3B

together with the spectrum of PSI-LHCI (Drop et al. 2011).

Figure 2. Fluorescence emission spectra at 77K of the thylakoids prepared from WT cells in state 1

(black) and state 2 (red). The spectra are normalized to their respective maximum.

Figure 3. PSI-LHCI-LHCII isolation. A: Sucrose density gradients for the separation of the PSI complexes. B: Absorption spectra at room temperature of some of the bands in A together with the spectrum of PSI-LHCI (Drop, Webber-Birungi et al. 2011). The spectra are normalized to the maximum (678-679 nm).

Page 106: University of Groningen From Photosystem I to Photosystem

105

It should be noted that the two lowest bands in all gradients have identical absorption

spectra, suggesting that they contain the same complexes but in a different aggregation

state. The spectra shown in Fig. 3B have a maximum at 678-9 nm and absorption above 700

nm, which clearly indicates that these bands all contain PSI complexes. However, the

complexes purified from both WT-His and Fl39 cells in state 2 have a higher chlorophyll b

content than either PSI-LHCI or the complex in state 1, as can be assessed by the difference

in absorption at 650 nm. This indicates that the PSI complexes purified from state 2 cells

have a larger antenna size than PSI-LHCI.

3.2 Photosystem II supercomplexes are stable in state 2

To check the effect of state transitions on the organization and composition of the

PSII supercomplexes, the flow-through of the Ni-column, which is strongly enriched in PSII

(Drop et al. 2014), was also loaded on sucrose gradients. In agreement with previous results

(Drop et al. 2014), nine bands plus some insolubilized material were visible in the gradients

in state 1 and 2 (Fig. 4A).

Figure 4. Isolation of PSII-LHCII supercomplexes from state 1 and 2. A: Fractionation of PSII sub/super-complexes from state 1 and state 2 cells by sucrose density gradient. B: Absorption spectra of fractions B3 (LHCII trimers) and B7-B9 containing the PSII-supercomplexes from state 2 membranes. The spectra are normalized to the maximum (676 nm).

A detailed analysis of the content of the individual bands was previously reported (Drop et

al, 2014). Fractions B7, B8 and B9 represent PSII supercomplexes with different antenna

sizes, as confirmed by the analysis of their absorption spectra (Fig. 4B). It is interesting to

Page 107: University of Groningen From Photosystem I to Photosystem

106

observe that fractions B7, B8 and B9 were present in both state 1 and state 2 gradients,

indicating that the supercomplexes exist in each state, although B8 and B9 were clearly less

intense in state 2 (minus 25% as determined by the analysis of the intensity in the

gradients), suggesting a lower stability (or a partial disassembly) of the PSII supercomplexes

in state 2.

3.3 PSI-LHCI-LHCII complexes contain all four LhcbM types and CP29

The protein composition of the PSI complexes purified from WT-His and Fl39 state 2 cells

(bands B3 and B4 in the respective gradients in figure 3A) was analyzed by SDS-PAGE

(Fig. 5A). The results confirm that both preparations contain PSI complexes, as indicated by

the presence of PsaA and -B (~65 kDa), Lhca (20-27 kDa) and small core subunits (10-20 kDa)

(Bassi, Soen et al. 1992; Ozawa, Onishi et al. 2010; Drop, Webber-Birungi et al. 2011).

Figure 5. Protein composition of the isolated PSI complexes from state 2 cells. A: SDS-PAGE showing the PSI preparation from state 2 of WT-His (B2) and Fl39 (B4) compared with purified Lhcb (fraction 2 from the gradient in figure 3A) and PSI-LHCI, purified as in Drop et al. 2011. B: Western blotting against CP26 and CP29. Lanes 1: PSII-LHCII; Lanes 2: PSI-LHCI-LHCII from WT-his; Lane 3: PSI-LHCI-LHCII from Fl39.

Page 108: University of Groningen From Photosystem I to Photosystem

107

In addition, both fractions show four bands corresponding to LhcbM proteins type II, I, III

and IV, from top to bottom) and bands at the height of CP29 and CP26. As expected, no PSII

core subunits were visible in the preparations. Contamination due to ATPase was present

but since this complex does not bind pigments, it does not interfere with the measurement

performed in this study. The presence of CP29 and CP26 in the PSI supercomplex was

confirmed by western blotting (Fig. 5B). The data show that both CP29 and CP26 are present

in the PSI-LHCI-LHCII supercomplexes, but in different amounts, with clearly more CP29 than

CP26. The amount of these subunits in PSI-LHCI-LHCII complexes can be roughly quantified

considering that each of them contains 9% of the Chls of the PSII supercomplex (13 Chls

each (Pan, Li et al. 2011) out of the 145 Chls of the supercomplex (Drop et al. 2014)). In the

PSI-LHCI-LHCII from WT-his cells (Fig. 5A) the amount of CP29, as determined by the

intensity of the band in the gels, was up to 50% of the amount in the PSII-LHCII, while the

values were around 10-20% for CP26. Considering that an equal amount of Chls was loaded

in each lane, this means that in the PSI-LHCI-LHCII supercomplex CP29 accounts for 4.5%

and CP26 for 1-2% of the total chlorophylls.

3.4 LhcbM-type I, II and IV but not CP29 and LhcbM-type III are phosphorylated when

associated with PSI

Phosphorylation and dephosphorylation are the central events of state transitions as they

trigger the movements of Lhcb complexes between PSII and PSI (Andersson, Akerlund et al.

1982; Kyle, Staehelin et al. 1983). To determine the phosphorylation state of the individual

proteins in state 1 and state 2, the proteins of PSI as well as those of the PSII-LHCII

supercomplexes purified from state 1 and state 2 were analyzed with a phosphoprotein gel

stain, Pro-Q Diamond. In Fig.6, phosphoproteins were first detected (Fig. 6B) followed by

total protein staining with Coomassie Blue (Fig. 6A). In the PSI-LHCI-LHCII supercomplexes

from state 2, the bands corresponding to LhcbM-type IV (LhcbM1) and type I were strongly

stained with Pro-Q stain, indicating that they are phosphorylated. Type II (LhcbM5) was also

visible upon Pro-Q staining, while virtually no signal was observed for type III LhcbM

complexes (LhcbM2/7) and CP29, suggesting that those complexes are not phosphorylated

when associated with PSI. In the PSI WT complex from state 1, the only stained band

corresponded to LhcbM1. Interestingly, thylakoids in state 2 showed phosphorylation of the

bands corresponding to CP29 and CP26, but these phosphorylated complexes were found

Page 109: University of Groningen From Photosystem I to Photosystem

108

mainly associated with the PSII-LHCII complex. Phosphorylation of the core proteins was

also visible in the PSII-LHCII complexes both in state 1 and in state 2. In addition, not only

did the gel band in PSII-LHCII corresponding to type I LhcbM contain phosphorylated

proteins, in agreement with previous results (Iwai, Takahashi et al. 2008), but also that

corresponding to type IV LhcbM. Type I LHCII was the only phosphorylated band in the

trimeric LHCII fraction (B3 in the gradient in figure 4A) in state 2.

Figure 6. Photospohrylation state of the proteins in PSI and PSII (super)-complexes in state 1 and 2. A: SDS-page stained with Coomassie Blue; B: the same gel as in A stained with Pro-Q Diamond, which reveals the phosphoproteins. Lane 1: PSI-LHCI obtained in state 1 upon solubilization with α-DM (Drop et al. 2011); Lane 2: PSI-LHCI-LHCII from Fl39 cells; Lane 3: PSI-LHCI-LHCII from WT-His cells; Lane 4: PSII-LHCII from state 2; Lane 5: PSII-LHCII from state 1; Lane 6:LHCII trimers from state 2; Lane 7: PSI from state 1 upon solubilization with digitonin; Lane 8: thylakoids in state 2; Lane 9: thylakoids in state 1.

Page 110: University of Groningen From Photosystem I to Photosystem

109

3.5 The Chlorophyll a/b ratio of PSI-LHCI-LHCII is 2.95 vs. 4.4 of PSI-LHCI

The pigment composition of the PSI-LHCI-LHCII complex is reported in Table 1. As compared

to PSI-LHCI, the data show a decrease of the chlorophyll a/b ratio in PSI-LHCI-LHCII (2.95 vs.

4.4) and a relative increase of the neoxanthin/loroxanthin content, in agreement with the

presence of LHCII. Using the pigment composition of the individual complexes, the number

of Lhcb subunits associated with PSI in PSI-LHCI-LHCII can roughly be calculated. Considering

that PSI coordinates 240 Chls - 196 Chl a and 44 Chl b (Drop, Webber-Birungi et al. 2014) -

and LHCII contains 14 Chls per monomer – 8 Chl a and 6 Chl b (Liu, Yan et al. 2004) -

a decrease in Chl a/b ratio from 4.4 to 2.95 corresponds to the addition of 7 monomeric

Lhcb antenna to the PSI-LHCI complex.

It is interesting to underline that the Chl a/b ratio of the PSI complex in state 1 purified upon

digitonin solubilization is 4.15, clearly lower than that of the complex obtained upon

solubilization with α-DM. Doing the same calculation reported above, this suggests that the

complex has one additional subunit associated with it as compared with the PSI-LHCI

complex purified in α-DM.

Table 1. Pigment composition of PSI supercomplexes. The values of individual carotenoids are normalized to 240 Chls (a+b) for PSI-LHCI and to 336 Chls (a+b) for PSI-LHCI-LHCII (see below).

Chl a/b Chl/car Loroxanthin/

Neoxanthin

Violaxanthin Lutein β-carotene

PSI-LHCI (αDM) 4.4 ± 0.1 4.5 ± 0.3 4.8 ± 0.6 11.5 ± 0.6 16.8 ± 1.0 20.4 ± 0.3

PSI-LHCI-LHCII 2.95 ± 0.05 4.7 ± 0.9 13.9 ± 0.8 16.7 ± 1.4 21.6 ± 2.7 17.9 ± 2.4

3.6 The Lhcb subunits associated with PSI in the PSI-LHCI-LHCII complex efficiently transfer

excitation energy to the PSI core

To determine the absorption spectrum of the Lhcb fraction associated with PSI, the spectra

of PSI-LHCI-LHCII and PSI-LHCI were compared after normalization to the Chl content.

The results are shown in Fig. 7A: the difference spectrum has the typical feature of Lhcb

complexes of C. reinhardtii, with a maximum at 672 nm and a clear Chl b shoulder at 650

nm, as can be seen by comparison with the spectrum of the Lhcb fraction from the sucrose

Page 111: University of Groningen From Photosystem I to Photosystem

110

gradient (B2 in Fig. 4A), confirming that the observed differences are due to the addition of

Lhcb to the PSI-LHCI complex.

Figure 7. Spectroscopic properties of the PSI-LHCI-LHCII complex (red) compared to those of PSI-LHCI

(black). A: Absorption spectrum at room temperature. The spectra are normalized to the Chl content

of the two complexes, the difference spectrum (PSI-LHCI-LHCII minus PSI-LHCI) is in blue. The

spectrum of the Lhcb fraction (fraction B2 from the gradient in figure 4A) is in green. B: 77K

fluorescence emission spectra of PSI supercomplexes excited at 475 nm. The spectra are normalized

to the maximum. C: Fluorescence at RT of PSI-LHCI-LHCII upon excitation at 440 nm, 475 nm and 500

nm. Excitation and emission bandwidths were set to 3 nm.

To check the integrity of PSI-LHCI-LHCII and the capacity of the Lhcb to transfer excitation

energy to the PSI core, the fluorescence spectra were measured at 77K (Fig. 7B) and 5°C

(Fig. 7C). At 5°C the overlap of the emission spectra after excitation at 440 nm, 475 nm and

500 nm, which preferentially excite Chl a, Chl b and carotenoids, indicates that there are no

free pigments in the preparation and that the Lhcbs transfer excitation energy to PSI. This is

confirmed by the 77K fluorescence measurements, which show the typical red emission of

PSI, and only a small shoulder at 680 nm. In conclusion, the data indicate that the Lhcbs are

functionally associated with PSI-LHCI, although the shoulder at 680 nm indicates that the

efficiency of energy transfer is not 100%.

Figure 8. Electron micrograph of isolated negatively stained C. reinhardtii PSI-LHCI-LHCII with two

trimers (green box) and one trimer (yellow box). Space bar is 100 nm.

Page 112: University of Groningen From Photosystem I to Photosystem

111

3.7 In addition to nine Lhcas, the largest PSI-LHCI-LHCII complex contains three Lhc trimers

and one monomer located on the PsaH/L side of the complex

The structural organization of PSI-LHCI-LHCII (from both WT-His and mutant, Fig. 8) was

determined by electron microscopy. Original images showed a sample containing large

rectangular-shaped particles (in the following indicated as the “large” particle), but also

a smaller pear-shaped particle present at a ratio of roughly 1 to 5 (Fig. 9) (in the following

indicated as the “small” particle). Both types were analyzed by single particle averaging

(Fig. 10).

Figure 9. Comparison of maps of PSI-LHCI-LHCII particles. A: 2D map of sum of the PSI-LHCI-LHCII

particle purified from the Fl39 mutant, as presented in Figure 7. B: 2D map of the PSI-LHCI-LHCII

particle purified from WT-his, processed independently. Scale bar for both frames is 10 nm.

The projection map of the most abundant, large PSI-LHCI-LHCII particle at 2 nm resolution

shows densities of two trimers and one monomer (Fig. 10A). Fitting the high-resolution

structures of its components into the projection map clearly supports this conclusion

(Fig. 10B). The small particle has a different shape than the large one, because it lacks most

of the area occupied in the large PSI particle by the lower LHCII trimer (Fig. 10C). Because of

the lower abundance of this particle in the preparation, and thus in the EM images, only

small numbers of projections could be processed and hence the map has a lower resolution

and the antenna complexes cannot be fitted unambiguously. But if we consider that the

upper right part of the small and large particles (upper with reference to the orientation of

the particles in Fig. 10) are the same, the former may contain CP29 and LHCII trimer at the

same site as the latter (Fig. 10D). Modelling also indicates that there is enough space left for

another protein (red question mark) and it is very tempting to consider the low-abundant

CP26 protein as a candidate. A further observation about the state transition particles is the

substantial increase in surface in comparison to PSI-LHCI (Fig. 10E), whereas in higher plants

the PSI-LHCI-LHCII complex contains only one of the trimers (Kouril, Zygadlo et al. 2005)

(Fig. 10F).

Page 113: University of Groningen From Photosystem I to Photosystem

112

Figure 9. Single particle averaging of PSI complexes in state 2. A: 2D map of the largest PSI-LHCI-LHCII particle from the Fl39 mutant, which is the sum of the best 3000 projections from a set of 12,000. B: Overlap of the map with wire models of the high resolution structures of plant PSI, the 9 LHCI antenna proteins (Drop et al, 2011), two LHCII trimers and CP29. C: Map of a smaller state transition complex, from 500 summed projections. D: Tentative assignment of densities of the smaller PSI complex with CP29 and one LHCII trimer attached. The red question mark indicates an area which is too small to accommodate a trimeric LHCII. It may be occupied by a monomeric Lhcb protein. E: The C. reinhardtii PSI-LHCI particle, analyzed previously (Drop et al, 2011). F: Plant state transition PSI-LHCI-LHCII particle of Arabidopsis thaliana (Kouril et al, 2005a). Scale bar for all frames is 10 nm.

4. Discussion

State transitions represent a key regulatory process of the light reactions in Chlamydomonas

reinhardtii. The process has been studied in detail, especially regarding the trigger for the

transitions (Horton and Black 1980; Allen, Bennett et al. 1981; Allen and Forsberg 2001), the

role of phosphorylation (Bennett 1980; Depege, Bellafiore et al. 2003; Bonardi, Pesaresi et

al. 2005; Lemeille, Turkina et al. 2010), and the interplay with non-photochemical quenching

(Allorent, Tokutsu et al. 2013) and cyclic electron transport (Iwai, Takizawa et al. 2010;

Takahashi, Clowez et al. 2013). However, the consequences of the transitions on the

Page 114: University of Groningen From Photosystem I to Photosystem

113

functional and structural organization of the supercomplexes have not been elucidated.

Here we have purified and analyzed PSI and PSII supercomplexes in the different transition

states with the aim of filling this gap.

4.1 How is PSI-LHCI-LHCII organized? What is its antenna size?

Our data consistently show that the largest purified complex contains two trimers and one

monomer in addition to the PSI-LHCI complex, and that the Lhcbs transfer excitation energy

to PSI. The particle is larger than that of plants where only one LHCII trimer was found to be

associated with PSI in state 2 (Kouril, Zygadlo et al. 2005). The modeling of the EM data

shows that in addition to a trimer situated in a similar position as in plants, a second trimer

is located close to PsaB and PsaI on the other side of PsaH/L, and a monomer is involved in

its binding to the PSI core. This monomeric complex is likely CP29, which is present in

stoichiometric amounts in PSI-LHCI-LHCII, and whose presence in the supercomplex has

been suggested to be necessary for binding of LHCII to PSI (Kargul, Turkina et al. 2005;

Tokutsu, Iwai et al. 2009). The finding of a second type of particle, at lower abundance, hints

to a rather complex situation in state transitions, where more than just one particle may be

involved. Whatever the further details and implications, the increase of the antenna of PSI

in state 2 is substantial (~100 Chls), but it should be seen in proportion because the PSI-LHCI

with 9 LHCI proteins is already large. In fact, the addition of one to two LHCII trimers plus at

least CP29 is close to proportional to the enlargement of the 4 LHCI antenna proteins of

higher plant PSI with just one LHCII trimer.

4.2 Which Lhcb subunits are associated with PSI in C. reinhardtii?

Previous analysis of the composition of PSI-LHCII has suggested that CP29, CP26 and

LhcbM5 are associated with PSI in state 2 (Takahashi, Iwai et al. 2006). It was also shown

that the extent of state transitions was reduced in the absence of LhcbM2/7, suggesting an

involvement of these complexes in state transitions (Ferrante, Ballottari et al. 2012),

although this was in contrast with the fact that these subunits are not phosphorylated

(Lemeille, Turkina et al. 2010). Our data show that all four LhcbM types are present in the

PSI-LHCI-LHCII complex together with CP29 and CP26. The data indicate that out of the 340

Chls of PSI-LHCI-LHCII, 4.5% belong to CP29 and 1-2% to CP26. Considering that both

complexes coordinate 13 Chls, it can be concluded that CP29 is present in 1:1 stoichiometry

with the core, while CP26 is present in a substoichiometric amount.

Page 115: University of Groningen From Photosystem I to Photosystem

114

The data show that of the LhcbM complexes associated with PSI, type IV, type I and type II

are phosphorylated, while type III is not. Therefore, it appears that phosphorylation of all

subunits is not necessary for their association with PSI in C.reinhardtii. Surprisingly, CP29 is

also not (or very little) phosphorylated when associated with PSI, even though

phosphorylated CP29 is found in the thylakoids in state 2 as previously reported (Lemeille,

Turkina et al. 2010).

4.3 PSII-LHCII supercomplexes are still present in state 2

The data show that PSII-LHCII can be purified from cells in state 2, although in a smaller

amount than in state 1, indicating that part of the supercomplexes are intact in those

conditions and that state transitions involve only a subset of Lhcb complexes (see below).

Interestingly, phosphorylated Lhcb, especially LhcbM type IV and I but also CP29 and CP26,

remain associated with the PSII supercomplexes in state 2, indicating that the

phosphorylation of the individual subunits does not necessary lead to supercomplex

disassembly, similar to what was observed in plants (Wientjes, Drop et al. 2013). As it has

been proposed that these complexes contain multiple phosophorylation sites (Hippler, Klein

et al. 2001; Turkina, Kargul et al. 2006; Iwai, Takahashi et al. 2008), it will be interesting to

compare the phosphorylation patterns of the subunits when associated with PSI or with

PSII.

4.4 Comparison of state transitions in C. reinhardtii and A. thaliana

State transitions have been largely studied in A. thaliana and C. reinhardtii, the two model

organisms of plants and green algae. It is generally accepted that the mechanism is the

same in the two organisms and that the main difference is the amount of LHCII participating

in the transitions: 10-15% in plants vs. 80% in C. reinhardtii (Delosme, Olive et al. 1996).

Here we clearly demonstrate that PSI-LHCI-LHCII in C. reinhardtii is indeed larger than in

higher plants, containing two LHCII trimers and one CP29 monomer, compared to only one

trimer in A. thaliana. However, this does not mean that the population of LHCII participating

in state transitions is larger in C. reinhardtii than in A. thaliana. We should take into account

the difference in PSI and PSII antenna size in the two organisms; we have recently shown

that the membranes of C. reinhardtii contain around seven LHCII trimers per monomeric PSII

core (Drop, Webber-Birungi et al. 2014), while this number varies from two to four in A.

thaliana (Kouril, Wientjes et al. 2012). It is the same in case of PSI, which is composed of

Page 116: University of Groningen From Photosystem I to Photosystem

115

four Lhca in plants and of nine in C. reinhardtii (Ben-Shem, Frolow et al. 2003; Drop,

Webber-Birungi et al. 2011). The stoichiometry P680: P700 is 1: 1 in this alga, meaning that

even if every PSI contains two LHCII trimers in state 2, the transition would involve less than

30% of the LHCII, while to match the proposed 80%, at least 5 LHCII trimers should associate

with PSI. Another difference between C. reinhardtii and A. thaliana is the involvement of

CP29 in the former but not in the latter. CP29 is part of the PSII-LHCII supercomplex and it is

located in between the core and the LHCII trimers, implying that its movement to PSI

requires the disassembly of the supercomplexes (Iwai, Takahashi et al. 2008). However,

here we show that part of the PSII-LHCII complexes is still present in state 2. As CP29 seems

to be necessary for the association of the LHCII trimer to PSI, this indicates that not all PSI

can contain two LHCII trimers, thus lowering the extent of the transition to values that

become closer to those of plants, in agreement with recent time-resolved data (Unlu, Drop

et al. 2014).

In summary, a new view on state transitions in C. reinhardtii is emerging from this

study. Most of the Lhcb complexes are shown to be able to act as antenna for both PSI and

PSII but the transition does not lead to the full disassembly of PSII-LHCII complex, suggesting

that only one subpopulation of Lhcb is involved in the transitions. It is also shown that

contrary to all expectations, several Lhcb subunits can associated with PSI in their non-

phosphorylated form, and to PSII in their phosphorylated form, indicating that the well-

accepted model in which phosphorylation leads to dissociation of Lhcb complexes from PSII

and to their association with PSI, cannot be applied to all subunits.

Acknowledgments

We thank Laura M. Roy for critical reading the manuscript. This work was supported by the

ERC Starting/Consolidator Grant 281341 (ASAP) to RC and by the FOM (Foundation for

fundamental research on matter) grant 10TBSC12-2 to EJB.

Page 117: University of Groningen From Photosystem I to Photosystem

116

References

Allen, J. F. (1992). "How Does Protein-Phosphorylation Regulate Photosynthesis." Trends in Biochemical Sciences 17(1): 12-17.

Allen, J. F., J. Bennett, et al. (1981). "Chloroplast protein phosphorylation couples plastoquinone redox state to distribution of excitation energy between photosystems." Nature 291: 25-29.

Allen, J. F. and J. Forsberg (2001). "Molecular recognition in thylakoid structure and function." Trends in Plant Science 6(7): 317-326.

Allorent, G., R. Tokutsu, et al. (2013). "A dual strategy to cope with high light in Chlamydomonas reinhardtii." Plant Cell 25(2): 545-557.

Andersson, B., H. E. Akerlund, et al. (1982). "Differential Phosphorylation of the Light-Harvesting Chlorophyll Protein Complex in Appressed and Non-Appressed Regions of the Thylakoid Membrane." Febs Letters 149(2): 181-185.

Bassi, R. (1985). "Spectral Properties and Polypeptide Composition of the Chlorophyll-Proteins from Thylakoids of Granal and Agranal Chloroplasts of Maize (Zea-Mays-L)." Carlsberg Research Communications 50(2): 127-143.

Bassi, R., S. Y. Soen, et al. (1992). "Characterization of Chlorophyll-a/b Proteins of Photosystem-I from Chlamydomonas-Reinhardtii." J.Biol.Chem. 267: 25714-25721.

Bellafiore, S., F. Barneche, et al. (2005). "State transitions and light adaptation require chloroplast thylakoid protein kinase STN7." Nature 433(7028): 892-895.

Ben-Shem, A., F. Frolow, et al. (2003). "Crystal structure of plant photosystem I." Nature 426(6967): 630-635.

Bennett, J. (1980). "Chloroplast Phosphoproteins - Evidence for a Thylakoid-Bound Phosphoprotein Phosphatase." European Journal of Biochemistry 104(1): 85-89.

Bennett, J., K. E. Steinback, et al. (1980). "Chloroplast Phosphoproteins - Regulation of Excitation-Energy Transfer by Phosphorylation of Thylakoid Membrane Polypeptides." Proceedings of the National Academy of Sciences of the United States of America-Biological Sciences 77(9): 5253-5257.

Boekema, E. J., H. van Roon, et al. (1999). "Supramolecular organization of photosystem II and its light-harvesting antenna in partially solubilized photosystem II membranes." European Journal of Biochemistry 266(2): 444-452.

Bonardi, V., P. Pesaresi, et al. (2005). "Photosystem II core phosphorylation and photosynthetic acclimation require two different protein kinases." Nature 437(7062): 1179-1182.

Bonaventura, C. and J. Myers (1969). "Fluorescence and oxygen evolution from Chlorella pyrenoidosa." Biochim.Biophys.Acta 189: 366-383.

Bulte, L., P. Gans, et al. (1990). "Atp Control on State Transitions Invivo in Chlamydomonas-Reinhardtii." Biochimica Et Biophysica Acta 1020(1): 72-80.

Chuartzman, S. G., R. Nevo, et al. (2008). "Thylakoid membrane remodeling during state transitions in Arabidopsis." Plant Cell 20(4): 1029-1039.

Croce, R., G. Canino, et al. (2002). "Chromophore organization in the higher-plant photosystem II antenna protein CP26." Biochemistry 41(23): 7334-7343.

Croce, R. and H. van Amerongen (2011). "Light-harvesting and structural organization of Photosystem II: from individual complexes to thylakoid membrane." J Photochem Photobiol B 104(1-2): 142-153.

Dekker, J. P. and E. J. Boekema (2005). "Supramolecular organization of thylakoid membrane proteins in green plants." Biochim.Biophys.Acta 1706: 12-39.

Delosme, R., J. Olive, et al. (1996). "Changes in light energy distribution upon state transitions: An in vivo photoacoustic study of the wild type and photosynthesis mutants from Chlamydomonas reinhardtii." Biochimica Et Biophysica Acta-Bioenergetics 1273(2): 150-158.

Page 118: University of Groningen From Photosystem I to Photosystem

117

Depege, N., S. Bellafiore, et al. (2003). "Rote of chloroplast protein kinase Stt7 in LHCII phosphorylation and state transition in Chlamydomonas." Science 299(5612): 1572-1575.

Drop, B., M. Webber-Birungi, et al. (2011). "Photosystem I of Chlamydomonas reinhardtii contains nine light-harvesting complexes (Lhca) located on one side of the core." J Biol Chem 286(52): 44878-44887.

Drop, B., M. Webber-Birungi, et al. (2014). "Light-harvesting complex II (LHCII) and its supramolecular organization in Chlamydomonas reinhardtii." Biochim Biophys Acta 1837(1): 63-72.

Elrad, D. and A. R. Grossman (2004). "A genome's-eye view of the light-harvesting polypeptides of Chlamydomonas reinhardtii." Current Genetics 45(2): 61-75.

Ferrante, P., M. Ballottari, et al. (2012). "LHCBM1 and LHCBM2/7 polypeptides, components of major LHCII complex, have distinct functional roles in photosynthetic antenna system of Chlamydomonas reinhardtii." Journal of Biological Chemistry 287(20): 16276-16288.

Galka, P., S. Santabarbara, et al. (2012). "Functional Analyses of the Plant Photosystem I-Light-Harvesting Complex II Supercomplex Reveal That Light-Harvesting Complex II Loosely Bound to Photosystem II Is a Very Efficient Antenna for Photosystem I in State II." Plant Cell 24(7): 2963-2978.

Germano, M., A. E. Yakushevska, et al. (2002). "Supramolecular organization of photosystem I and light- harvesting complex I in Chlamydomonas reinhardtii." FEBS Letters 525(1-3): 121-125.

Gorman, D. S. and R. P. Levine (1965). "Cytochrome f and plastocyanin: their sequence in the photosynthetic electron transport chain of Chlamydomonas reinhardi." Proc Natl Acad Sci U S A 54(6): 1665-1669.

Gulis, G., K. V. Narasimhulu, et al. (2008). "Purification of His(6)-tagged photosystem i from Chlamydomonas reinhardtii." Photosynthesis Research 96(1): 51-60.

Hippler, M., J. Klein, et al. (2001). "Towards functional proteomics of membrane protein complexes: analysis of thylakoid membranes from Chlamydomonas reinhardtii." Plant J 28(5): 595-606.

Horton, P. and M. T. Black (1980). "Activation of Adenosine 5'-Triphosphate-Induced Quenching of Chlorophyll Fluorescence by Reduced Plastoquinone - the Basis of State-I-State-Ii Transitions in Chloroplasts." Febs Letters 119(1): 141-144.

Iwai, M., Y. Takahashi, et al. (2008). "Molecular remodeling of photosystem II during state transitions in Chlamydomonas reinhardtii." Plant Cell 20(8): 2177-2189.

Iwai, M., K. Takizawa, et al. (2010). "Isolation of the elusive supercomplex that drives cyclic electron flow in photosynthesis." Nature 464(7292): 1210-1213.

Jansson, S. (1999). "A guide to the Lhc genes and their relatives in Arabidopsis." Trends Plant Sci. 4: 236-240.

Kargul, J., J. Nield, et al. (2003). "Three-dimensional reconstruction of a light-harvesting complex I-photosystem I (LHCI-PSI) supercomplex from the green alga Chlamydomonas reinhardtii - Insights into light harvesting for PSI." Journal of Biological Chemistry 278(18): 16135-16141.

Kargul, J., M. V. Turkina, et al. (2005). "Light-harvesting complex II protein CP29 binds to photosystem I of Chlamydomonas reinhardtii under State 2 conditions." Febs Journal 272(18): 4797-4806.

Kouril, R., E. Wientjes, et al. (2012). "High-light vs. low-light: Effect of light acclimation on photosystem II composition and organization in Arabidopsis thaliana." Biochim Biophys Acta 1827(3): 411-419.

Kouril, R., A. Zygadlo, et al. (2005). "Structural characterization of a complex of photosystem I and light-harvesting complex II of Arabidopsis thaliana." Biochemistry 44(33): 10935-10940.

Kouril, R., A. Zygadlo, et al. (2005). "Supercomplexes of photosystem I and antenna proteins in green plants and cyanobacteria." Febs Journal 272: 449-449.

Kyle, D. J., L. A. Staehelin, et al. (1983). "Lateral Mobility of the Light-Harvesting Complex in Chloroplast Membranes Controls Excitation-Energy Distribution in Higher-Plants." Archives of Biochemistry and Biophysics 222(2): 527-541.

Page 119: University of Groningen From Photosystem I to Photosystem

118

Lemeille, S., M. V. Turkina, et al. (2010). "Stt7-dependent Phosphorylation during State Transitions in the Green Alga Chlamydomonas reinhardtii." Molecular & Cellular Proteomics 9(6): 1281-1295.

Lemeille, S., A. Willig, et al. (2009). "Analysis of the Chloroplast Protein Kinase Stt7 during State Transitions." Plos Biology 7(3): 664-675.

Liu, Z., H. Yan, et al. (2004). "Crystal structure of spinach major light-harvesting complex at 2.72 A resolution." Nature 428(6980): 287-292.

Lunde, C., P. E. Jensen, et al. (2000). "The PSI-H subunit of photosystem I is essential for state transitions in plant photosynthesis." Nature 408(6812): 613-615.

Maroc, J., J. Garnier, et al. (1989). "Chlorophyll-Protein Complexes Related to Photosystem-I in Chlamydomonas-Reinhardtii." Journal of Photochemistry and Photobiology B-Biology 4(1): 97-109.

Merchant, S. S., S. E. Prochnik, et al. (2007). "The Chlamydomonas genome reveals the evolution of key animal and plant functions." Science 318(5848): 245-250.

Minagawa, J. (2011). "State transitions-The molecular remodeling of photosynthetic supercomplexes that controls energy flow in the chloroplast." Biochimica Et Biophysica Acta-Bioenergetics 1807(8): 897-905.

Murata, N. (1969). "Control of excitation transfer in photosynthesis. I. Light- induced change of chlorophyll a fluorescence in Porphyridium cruentum." Biochim.Biophys.Acta 172: 242-251.

Nelson, N. (2009). "Plant Photosystem I - The Most Efficient Nano-Photochemical Machine." Journal of Nanoscience and Nanotechnology 9(3): 1709-1713.

Oostergetel, G. T., W. Keegstra, et al. (1998). "Automation of specimen selection and data acquisition for protein electron crystallography." Ultramicroscopy 74(1-2): 47-59.

Ozawa, S., T. Onishi, et al. (2010). "Identification and characterization of an assembly intermediate subcomplex of photosystem I in the green alga Chlamydomonas reinhardtii." Journal of Biological Chemistry 285(26): 20072-20079.

Pan, X., M. Li, et al. (2011). "Structural insights into energy regulation of light-harvesting complex CP29 from spinach." Nat Struct Mol Biol 18(3): 309-315.

Porra, R. J., W. A. Thompson, et al. (1989). "Determination of Accurate Extinction Coefficients and Simultaneous-Equations for Assaying Chlorophyll-a and Chlorophyll-B Extracted with 4 Different Solvents - Verification of the Concentration of Chlorophyll Standards by Atomic-Absorption Spectroscopy." Biochimica Et Biophysica Acta 975(3): 384-394.

Pribil, M., P. Pesaresi, et al. (2010). "Role of Plastid Protein Phosphatase TAP38 in LHCII Dephosphorylation and Thylakoid Electron Flow." Plos Biology 8(1).

Rebeille, F. and P. Gans (1988). "Interaction between Chloroplasts and Mitochondria in Microalgae - Role of Glycolysis." Plant Physiology 88(4): 973-975.

Rochaix, J. D. (2007). "Role of thylakoid protein kinases in photosynthetic acclimation." Febs Letters 581(15): 2768-2775.

Shapiguzov, A., B. Ingelsson, et al. (2010). "The PPH1 phosphatase is specifically involved in LHCII dephosphorylation and state transitions in Arabidopsis." Proc Natl Acad Sci U S A 107(10): 4782-4787.

Standfuss, R., A. C. T. van Scheltinga, et al. (2005). "Mechanisms of photoprotection and nonphotochemical quenching in pea light-harvesting complex at 2.5A resolution." Embo Journal 24(5): 919-928.

Stauber, E. J., A. Busch, et al. (2009). "Proteotypic profiling of LHCI from Chlamydomonas reinhardtii provides new insights into structure and function of the complex." Proteomics 9(2): 398-408.

Takahashi, H., S. Clowez, et al. (2013). "Cyclic electron flow is redox-controlled but independent of state transition." Nat Commun 4:(1954): 1-8.

Page 120: University of Groningen From Photosystem I to Photosystem

119

Takahashi, H., M. Iwai, et al. (2006). "Identification of the mobile light-harvesting complex II polypeptides for state transitions in Chlamydomonas reinhardtii." Proc Natl Acad Sci U S A 103(2): 477-482.

Tikkanen, M., M. Pippo, et al. (2006). "State transitions revisited - a buffering system for dynamic low light acclimation of Arabidopsis." Plant Molecular Biology 62(4-5): 779-793.

Tokutsu, R., M. Iwai, et al. (2009). "CP29, a monomeric light-harvesting complex II protein, is essential for state transitions in Chlamydomonas reinhardtii." J Biol Chem 284(12): 7777-7782.

Turkina, M. V., J. Kargul, et al. (2006). "Environmentally modulated phosphoproteome of photosynthetic membranes in the green alga Chlamydomonas reinhardtii." Mol Cell Proteomics 5(8): 1412-1425.

Unlu, C., B. Drop, et al. (2014). "State transitions in Chlamydomonas reinhardtii strongly modulate the functional size of photosystem II but not of photosystem I." Proc Natl Acad Sci U S A.

Wientjes, E., B. Drop, et al. (2013). "During state 1 to state 2 transition in Arabidopsis thaliana the Photosystem II supercomplex gets phosphorylated but does not disassemble." J Biol Chem 288(46), 32821–32826.

Wientjes, E., H. van Amerongen, et al. (2013). "LHCII is an antenna of both photosystems after long-term acclimation." Biochim Biophys Acta 1827(3): 420-426.

Zito, F., G. Finazzi, et al. (1999). "The Qo site of cytochrome b(6)f complexes controls the activation of the LHCII kinase." Embo Journal 18(11): 2961-2969.

Page 121: University of Groningen From Photosystem I to Photosystem

120

Page 122: University of Groningen From Photosystem I to Photosystem

121

CHAPTER 5

During state 1 to state 2 transition in Arabidopsis thaliana the

Photosystem II supercomplex gets phosphorylated but does not

disassemble

Emilie Wientjes,1 Bartlomiej Drop1, Roman Kouřil,2,3 Egbert J. Boekema,2 and Roberta Croce1

1Department of Physics and Astronomy, Faculty of Sciences, VU University Amsterdam, 1081 HV Amsterdam, The Netherlands,

2Groningen Biomolecular Sciences & Biotechnology Institute, Faculty of Mathematics and Natural Sciences, University of Groningen, 9747 AG, Groningen, The Netherlands,

3Centre of the Region Haná for Biotechnological and Agricultural Research, Department of Biophysics, Faculty of Science, Palacký University, 783 71 Olomouc, Czech Republic.

Based on Journal of Biological Chemistry 2013, 288(46): 32821-32826

Page 123: University of Groningen From Photosystem I to Photosystem

122

Abstract

Plants are exposed to continuous changes in light quality and quantity that challenge

the performance of the photosynthetic apparatus and have evolved a series of mechanisms

to face this challenge. In this work we have studied state transitions, the process that

redistributes the excitation pressure between Photosystems I and II (PSI/PSII) by the

reversible association of LHCII, the major antenna complex of higher plants, with either one

of them upon phosphorylation/ dephosphorylation. By combining biochemical analysis and

electron microscopy we have studied the effect of state transitions on the composition and

organization of Photosystem II in Arabidopsis thaliana. Two LHCII trimers (called trimers M

and S) are part of the PSII supercomplex, while up to two more are loosely associated with

PSII in state 1 in higher plants (called “extra” trimers). Here we show that the LHCII from the

“extra” pool migrates to PSI in state 2, thus leaving the PSII supercomplex and the semi-

crystalline PSII arrays intact. In state 2 not only is the mobile LHCII phosphorylated, but also

the LHCII in the PSII supercomplexes. This demonstrates that PSII phosphorylation is not

sufficient for disconnecting LHCII trimers S and M from PSII and for their migration to PSI.

Page 124: University of Groningen From Photosystem I to Photosystem

123

1. Introduction

Oxygenic photosynthetic organisms harvest light via two large pigment-protein assemblies

called photosystem I (PSI) and II (PSII). The two photosystems, connected through the

plastoquinol pool and the cytochrome b6f complex, work in series to transform light energy

into chemical energy (Nelson and Yocum 2006). Both photosystems have a specific

absorption spectrum. In plants and green algae, PSII is rich in Chlorophyll b (absorption

maxima around 475nm and 650nm) which is coordinated by light- harvesting complexes,

while PSI is rich in Chlorophyll a (absorption maxima around 440nm and 680nm). In

addition, PSI has Chlorophylls that can harvest light at wavelengths longer than 700nm

(Croce, Zucchelli et al. 1996). For optimal performance of linear electron transport the

excitation pressure on the two photosystems needs to be balanced. In natural fluctuating-

light environments, a mechanism known as state transitions serves to redistribute the

excitation energy between PSI and PSII (Bonaventaru and Myers 1969; Murata 1969). In

higher plants and green algae, this is achieved by the migration of part of the major light

harvesting complex (LHCII) between PSI and PSII. State 1 is traditionally defined as the

condition in which PSI is preferentially excited and all LHCII is associated with PSII. When

light conditions change, favoring PSII excitation, a mobile pool of LHCII moves from PSII and

associates with PSI (Allen 1992). This process is triggered by phosphorylation of LHCII by the

Stn7 kinase, which is activated upon oxidative reduction of the plastoquinol pool (Allen,

Bennett et al. 1981; Larsson, Jergil et al. 1983). Conversely, in state 2 preferential excitation

of PSI oxidizes the plastoquinol pool inactivating the Stn7 kinase. The constitutively active

LHCII phosphatase, (TAP38 (Pribil, Pesaresi et al. 2010)/ PPH1 (Shapiguzov, Ingelsson et al.

2010)), dephosphorylates the mobile LHCII which moves to PSII. Recently we showed that

LHCII is associated with both photosystems in most natural light conditions and it

disconnects from PSI only upon far-red illumination or high light stress, this is contrary to

the common assumption that LHCII is only associated to PSI as short term response

(Wientjes, van Amerongen et al. 2013). If exposed to different light qualities for a long time,

plants can adjust their PSI to PSII stoichiometry (Chow, Melis et al. 1990; Melis 1991; Allen

and Pfannschmidt 2000; Allen, Santabarbara et al. 2011; Hogewoning, Wientjes et al. 2012).

Instead, long term changes in light intensity lead to changes in the antenna size of both PSI

and PSII by regulating the level of the major light harvesting complex (LHCII) which functions

as an antenna for both photosystems (Wientjes, van Amerongen et al. 2013).

Page 125: University of Groningen From Photosystem I to Photosystem

124

PSII is composed of a core complex, where the primary photochemistry takes place, and a

peripheral antenna system, encoded by the Lhcb1-6 genes (Jansson 1999). LHCII is present

as a trimer composed of a combination of the Lhcb1-3 gene products. A threonine in the N-

terminus of Lhcb1 and Lhcb2 is the target for Stn7 (Bennett 1991). The minor Lhcbs consist

of three monomers, Lhcb4-6, also named CP29, CP26 and CP24. Lhcbs associate with

dimeric PSII cores to form PSII supercomplexes. The largest PSII supercomplex observed in

A.thaliana contains one copy of each minor complex and two LHCII trimers per core (C),

which are indicated as S (strongly bound) and M (moderately bound) forming the C2S2M2

supercomplex. Several PSII supercomplexes can associate to form megacomplexes (Dekker

and Boekema 2005). Depending on light growth conditions up to two more LHCII trimers per

PSII core can be present in the thylakoids (Peter and Thornber 1991; Dekker and Boekema

2005; Kouril, Wientjes et al. 2013), but their physical interaction with the PSII supercomplex

is rather weak or possibly absent in the sense that part of these complexes are probably not

in direct contact with the core, although they are still able to transfer energy to it via other

antennas (Wientjes, van Amerongen et al. 2013). We call these trimers “extra” LHCII

(Wientjes, van Amerongen et al. 2013). The grana membrane in St1 contains the PSII

supercomplexes (mainly in the C2S2M2 configuration) and “extra” LHCII which are located in

between the supercomplexes but probably also at the grana margins (Kouril, Wientjes et al.

2013).

Regarding the mobile LHCII involved in state transitions, it is known that the association of

LHCII with PSI occurs on the PsaH and PsaL side (Lunde, Jensen et al. 2000) and involves one

LHCII trimer (Kouril, Zygadlo et al. 2005). However, less is known about the association of

this mobile LHCII with PSII in St1. Although several models have been proposed of how state

transitions influence PSII organization, the identity of the trimer that moves from PSII to PSI

is still under debate. The involvement of trimer-M (and even S) has been suggested in

various models (Dietzel, Braautigam et al. 2011; Minagawa 2011) but it is controversially

discussed in literature. On the one hand Lhcb3, which is a component of trimer-M (Bassi and

Dainese 1992; Hankamer, Nield et al. 1997), was not found in the stroma lamellae in state 2

(Bassi, Giacometti et al. 1988) or in the PSI-LHCII supercomplex (Galka, Santabarbara et al.

2012; Wientjes, van Amerongen et al. 2013). On the other hand, upon state 1 -> state 2

transition an increased fraction of C2S2M1 and a decreased fraction of C2S2M2 complexes

was observed, suggesting that trimer-M goes from PSII to PSI (Kouril, Zygadlo et al. 2005).

Page 126: University of Groningen From Photosystem I to Photosystem

125

This was further supported by the finding that the rate of state transitions was enhanced in

CP24 and Lhcb3 KO mutants, where trimer-M is not (KO CP24) or less (KO Lhcb3) stably

associated with the PSII supercomplex (Kovacs, Damkjaer et al. 2006; Damkjaer, Kereiche et

al. 2009). More recently it was proposed that remodeling of PSII supercomplexes is required

for state transitions (2011), as has been suggested before for the green alga

Chlamydomonas reinhardtii (Iwai, Takahashi et al. 2008), while Tikkanen et al (2008)

propose that both PSI and PSII move to the grana margins in state 2 where they share LHCII.

In this study we combine biochemical and electron microscopy analysis to investigate the

effect of state transitions on the PSII supercomplex organization and to determine which

LHCII trimer(s) is/are involved in this process.

2. Materials and Methods

2.1 Plant Material

A.thaliana (Col) WT plants were grown at 100 μE/m2/s, 70% humidity, 22oC, and 8h/16h

day/night regime (Plant Climatics Percival Growth Chamber, Model AR-36L, Germany). Only

leaves fully exposed to the light were used for thylakoid isolation.

2.2 Plant treatment and thylakoid isolation

Leaves were harvested either directly after overnight dark adaptation (state 1) or after

50min of 20 μE/m2/s white light treatment (state 2). Harvested leaves were directly

transferred to ice/water. Thylakoid membranes were prepared according to (Caffarri, Kouril

et al. 2009) with the addition of 10mM of NaF to all buffers, to inhibit phosphatase activity.

The membrane was resuspended in 20mM Hepes pH 7.5, 0.4M sorbitol, 15mM NaCl, 5mM

MgCl2 and 10mM NaF, quickly frozen in N2 (l) and stored at 193 K until use. The Chlorophyll

a/b ratio (measured as in (Croce, Canino et al. 2002)) was the same for St1 and St2

thylakoids, indicating that no changes in PSI/PSII ratio had occurred during the 50min of

light treatment.

2.3 77K fluorescence measurements

The 77 K emission spectra were measured in 66% w/v glycerol, 10 mM Hepes pH 7.5, 5 mM

MgCl2, 15 mM NaCl and 10 mM NaF with a Fluorolog (Jobin-Yvon) with 3.5 nm bandwidth in

excitation and 2 in emission. The OD680 of the sample was <0.1 and the diameter of the

circular cuvette was 4mm.

Page 127: University of Groningen From Photosystem I to Photosystem

126

2.4 Polyacrylamide gel electrophoresis (PAGE)

First or second dimension denaturing PAGE was performed with the Tris-Tricine system

(Schagger 2006) at a 14.5% acrylamide concentration. Blue native PAGE was performed as in

(Jarvi, Suorsa et al. 2011), with the 25BTH20G buffer and an acrylamide/bisacrylamide ratio

of 32:1 in both stacking (3.5%) and resolving (4 to 14%) gel. The final Chl concentration was

0.5mg/ml and the final detergent concentration was 1% α- or -n-Dodecyl-D-maltoside

(α/ -DM) or 1% digitonin/0.1% α-DM as used in (Galka, Santabarbara et al. 2012). It should

be noted that the very small amount of α-DM added to digitonin does not have any effect

in the solubilization of the membrane but only helps the solubilization of digitonin (Galka,

Santabarbara et al. 2012). The cathode buffer was supplemented with 0.02% Coomassie

Blue G (Serva). Second dimension PAGE was performed as in (Jarvi, Suorsa et al. 2011). Pro-

Q Diamond Phosphoprotein Gel Stain (Molecular Probes) was used as described in the user

manual with fluorescence detection, except that the samples were not desalted and

delipidated. We confirmed that the signal was linear with the protein concentration. Gels

were stained with 0.025% Coomassie Blue G in 10% acetic acid as described in (Schagger

2006). The Coomassie and Pro-Q Diamond stained gels were imaged with ImageQuant LAS-

4000 (GE Healthcare).

2.5 Electron Microscopy (EM) and image analysis

Thylakoid membranes were solubilized in 20 mM Bis-Tris buffer pH 6.5, 5 mM MgCl2 using

digitonin (0.5 mg of chls /ml, 0.5% digitonin) for 20 min at 4 °C with slow stirring, followed

by centrifugation in an Eppendorf table centrifuge for 10 minutes. The pellet, which

contained the non-solubilized grana thylakoid membranes, was washed one more time

using the same buffer, centrifuged for 5 minutes and then used for EM analysis. Specimens

were prepared by negative staining with 2% uranyl acetate on glow-discharged carbon-

coated copper grids. Electron microscopy was performed as in (Caffarri, Kouril et al. 2009)

with 80000x magnification and semi-automated data acquisition (Oostergetel, Keegstra et

al. 1998). Sub-areas (256x256 pixels) of PSII arrays selected from individual electron

micrographs were analyzed separately. To determine a density of PSII complexes in

a membrane area mid-mass positions of the PSII core complexes were marked manually

using Groningen Image Processing software (Grip).

Page 128: University of Groningen From Photosystem I to Photosystem

127

The overlay of the C2S2M2 projection maps on the micrographs of the grana membranes

with randomly distributed PSII complexes was based on a cross-correlation alignment using

GRIP software. To enhance a contrast of PSII complexes and to facilitate their localization in

the grana membrane high frequency noise was filter out.

3. Results

3.1 State transitions

In this study the plants were brought to state 1 by dark adaptation, while state 2 was

induced by 50 minutes of low intensity white light treatment, conditions that are

physiological for the plant. To check if this treatment was successful, the level of LHCII

phosphorylation in the thylakoids was evaluated by Pro-Q Diamond Phosphorylation gel

stain (Fig. 1A). As expected LHCII was far less phosphorylated after dark adaptation (state 1)

then after light adaptation (state 2). Furthermore, the 77K emission spectrum shows

enhanced 735nm (PSI 77K fluorescence) emission in state 2 thylakoids compared to those in

state 1, confirming that state 1 and state 2 were successfully induced.

Figure 1. Analysis of state 1 (St1) and state 2 (St2) thylakoid analysis. A: Coomassie blue (Coom) and Pro-Q Diamond Phosphorylation (Pro-Q) stained SDS-PAGE gel. Gels were loaded with St1 or St2 thylakoids (1 μg on Chl basis). B: 77K fluorescence of St1 and St2 thylakoids. Excitation was at 484nm, normalized to the maximum in the 690-700nm PSII core emission region.

3.2 Influence of state transition on the organization of PSII supercomplexes

To compare the organization of PSII supercomplexes in state 1 and state 2, the thylakoid

membranes in the two states were solubilized with β-DM or with the milder α-DM, and

Page 129: University of Groningen From Photosystem I to Photosystem

128

loaded on a blue native gel (Fig.2). Freshly prepared thylakoids where used for this

experiment to avoid freeze/thaw cycles that might affect membrane and supercomplex

solubilisation. PSII supercomplexes were present in equal amount in state 1 and state 2 for

both solubilization conditions, indicating that the amount of the supercomplexes in intact

state 1 and state 2 membranes is the same. This idea is further supported by using a very

mild solubilisation condition (1% digitonin, 0.1% α-DM, Fig. 2) that solubilizes only the grana

margins, which is the region where the largest difference in the supercomplexes

organization between state 1 and state 2 can be expected. Also in this case the amount of

supercomplexes were identical in state 1 and state 2, indicating that not even the

supercomplexes in the grana margin disassemble during state transitions.

Figure 2. Blue native PAGE of thylakoids isolated from plants in state 1 (St1) or state 2 (St2). Solubilization conditions are indicated.

It should be added that when not using fresh thylakoids, the amount of supercomplexes

decreases in state 2 as compared to state 1, especially upon treatment with the β-DM (data

not shown), suggesting that the supercomplexes in state 2 are less stable than those in

state 1.

3.3. Organisation of PSII supercomplexes in the membrane

The organization of PSII supercomplexes in the grana membrane was further investigated by

electron microscopy (EM). The analysis reveals pairs of grana membranes where the typical

Page 130: University of Groningen From Photosystem I to Photosystem

129

features of PSII core complex are observed (Fig. 3). The major part (>90%) of the inspected

micrographs (555 for state 1 and 480 for state 2) showed a random organization of PSII

supercomplexes in the membrane (Fig. 3 A,D), while the other fraction (about 7% in St1 and

9% in state 2) showed a semi-crystalline organization (Fig. 3 B,E). The very similar amount of

crystalline arrays in the two states is not in agreement with the recent model of Dietzel et al.

(Dietzel, Braautigam et al. 2011), which suggests that disassembly of semi-crystalline PSII

arrays is a prerequisite for the state 1 -> state 2 transition. To identify the type of PSII

supercomplexes present in the semi-crystalline arrays in the two states, image analysis of

sub-areas of 2D arrays followed by fitting of the projection map to an X-ray pseudo-atomic

model of the PSII supercomplex (Caffarri, Kouril et al. 2009) was performed. In both state 1

and state 2, all supercomplexes in the 2D arrays were of the C2S2M2 type (Fig. 3C,F).

To investigate the effect of state transitions on the organization of PSII in the random

regions, the PSII density in the random area was determined. The white spot in the

micrographs represent PSII core complexes which protrude out of the membrane. About

1200 PSII particles were manually selected from 8 micrographs for both state 1 and state 2.

The average PSII density was 1549 ± 128 particles/μm2 (data are presented as mean ± S.D.,

n = 8 micrographs) in state 1 and 1601 ± 146 particles/μm2 in state 2, meaning that the PSII

density does not significantly change during state transitions (there is no statistically

significant difference at P=0.05 level determined using the Student´s test), while differences

were observed with the same method when the amount of LHCII changes (Kouril, Wientjes

et al. 2013). To give an idea of the possible organization of the supercomplexes in the

membrane, the position of the PSII complexes in state 1 and state 2 membranes were

highlighted by overlaying the projection map with the C2S2M2 supercomplex (Caffarri, Kouril

et al. 2009) (see Methods, Fig. 3B and 3D, respectively). The pattern is very similar in both

state 1 and state 2 indicating (i) a very similar supercomplex distribution and (ii) that the

large majority of PSII complexes can be present in the form of the C2S2M2 supercomplex, in

agreement with previous results obtained with cryo-electron microscopy (Kouril,

Oostergetel et al. 2011). We would like to stress here that it was not possible to investigate

the PSII density in the grana-margins. A change in the PSII density could occur in this region.

In conclusion, the molecular and supramolecular organization of PSII supercomplexes does

not change during state transitions; neither the occurrence/ composition of the crystalline

Page 131: University of Groningen From Photosystem I to Photosystem

130

arrays (from which the megacomplexes originate), nor the grana region with randomly

oriented supercomplexes were altered.

Figure 3. Organization of PSII complexes in the thylakoid membrane. Examples of electron micrographs of negatively stained pairs of thylakoid membranes with either a random (state 1(St1) A, state 2 (St2) C) or ordered (St1 E,St2 F) organization of PSII complexes. To give an idea of the possible organization of the supercomplexes in the membrane, in (B, D) a subset of PSII complexes in the membrane are highlighted by the overlay of the projection maps of C2S2M2 particles. Insets in (E, F): single particle analysis of sub-areas of ordered arrays of PSII supercomplexes selected from the particular electron micrograph. Averaged projections of 2D arrays of PSII complexes were assigned by fitting the structural pseudo-atomic model of the PSII C2S2M2 supercomplex according to (Caffarri, Kouril et al. 2009) (PSII core complex, the minor antenna CP29 and S-type of light-harvesting trimer are depicted in light green, M-type of light harvesting trimer, minor antennae CP24 and CP26 are depicted in light blue, dark salmon and light pink, respectively) (St1 E, St2 F).

3.4 Phosphorylation state of PSII supercomplexes

To clarify which LHCII complexes are phosphorylated in the different states, the thylakoid

membranes of plants in state 1 and state 2 were separated by blue-native PAGE and their

Page 132: University of Groningen From Photosystem I to Photosystem

131

protein content analyzed in the second dimension by SDS-PAGE. The phosphorylated

proteins were stained with Pro-Q Diamond stain, after which Coomassie stain was used to

visualize the total protein content (Fig. 4).

Figure 4. Phosphorylation state of PSII supercomplexes. State1 and state 2 thylakoids were solubilized with 1% α-DM and the supercomplexes were separated on blue native PAGE (BN, top) and the protein on SDS-PAGE. Total protein content was stained with Coomassie Blue (middle), and phosphorylated proteins were visualized with Diamond Pro-Q Phosophostain (bottom). Identities of PS supercomplexes (top, PSII super is PSII supercomplex, M-PSII is monomeric PSII core), important PSII proteins (middle) and phosphorylated PSII proteins (bottom) are indicated.

The PSI-LHCII complex is not present in this gel, as it disassembles into PSI and LHCII trimers

upon α-DM solubilization. The analysis shows that Lhcb1,2 present in the state 2 LHCII

trimers are phosphorylated. Indeed it has been shown previously that the LHCII

polypeptides associated with PSI are phosphorylated (Galka, Santabarbara et al. 2012;

Wientjes, van Amerongen et al. 2013), in agreement with the standard model that LHCII

moves to PSI when phosphorylated and to PSII upon dephosphorylation (Allen 1992;

Bellafiore, Bameche et al. 2005; Pribil, Pesaresi et al. 2010; Shapiguzov, Ingelsson et al.

2010). However, in contradiction with the general model, the gel shows that intact PSII

supercomplexes are also phosphorylated in state 2, confirming recent results of Grieco et al.

Page 133: University of Groningen From Photosystem I to Photosystem

132

(Grieco, Tikkanen et al. 2012). Thus, phosphorylated LHCII in the PSII supercomplex does not

move to PSI, which means that the standard state transitions model needs to be revised.

4. Discussion

4.1 The mobile pool of LHCII involved in state transitions consists of the “extra” (non-PSII

supercomplex) LHCII

In A. thaliana two LHCII trimers (S and M) per monomeric PSII core are located in the PSII

supercomplex (Dekker and Boekema 2005; Caffarri, Kouril et al. 2009), while up to two

“extra” LHCIIs per PSII monomer are present in the thylakoid membrane (Peter and

Thornber 1991; Kouril, Wientjes et al. 2013). Our investigation of their role in state

transitions shows that the involvement of trimer-M and S can be excluded: the absence of

trimer S would lead to the full disassembly of PSII supercomplexes, which does not occur as

we demonstrate here. Mobility of trimer-M can also be excluded, as Lhcb3, a marker for

this trimer, is not a component of the mobile LHCII fraction (Galka, Santabarbara et al. 2012;

Wientjes, van Amerongen et al. 2013). These findings lead to the conclusion that the “extra”

LHCIIs are the mobile antenna which can migrate between PSI and PSII, in agreement with

the proposal of Kyle et al. (Kyle, Staehelin et al. 1983). It is likely that these “extra” LHCIIs

are located near the grana-margins in state 1, such that transfer into the stroma lamellae

(state 2) only needs a short migration distance. This might explain why the PSII density

inside the PSII grana does not change upon state transitions (Fig. 3) and (Kyle, Staehelin et

al. 1983)).

The migration of the “extra” LHCIIs has clear advantages for the plant. Inside the PSII

supercomplex, the short distance between the trimers and the PSII core allows for fast

excitation-energy transfer and thus high PSII operating efficiency (Broess, Trinkunas et al.

2008). The “extra” LHCIIs are located more distantly from the PSII core, which leads to

slower, and thus less efficient, excitation-energy migration to the core (Wientjes, van

Amerongen et al. 2013). Thus, evolution has “chosen” to send the least efficient LHCII

antenna to PSI, while the efficient PSII supercomplex remains fully functional.

4.2 Phosphorylated LHCII does not have to move to PSI

According to the accepted model of state transitions, the phosphorylation of LHCII induces

its dissociation from PSII and its movement/association to PSI. Only in mutants lacking PSI

Page 134: University of Groningen From Photosystem I to Photosystem

133

(Delosme, Olive et al. 1996) or the LHCII anchor point in PSI (Lunde, Jensen et al. 2000) was

it observed that phosphorylated LHCII could not dock to PSI and stayed in the grana,

supporting the “molecular recognition model” (Lunde, Jensen et al. 2000; Allen and

Forsberg 2001). However, this observation has several possible explanations:

(i) phosphorylation involves only the “extra” LHCII, which in the absence of PSI cannot move

to PSI; (ii) phosophorylation involves also trimer M and S (the trimers that are part of the

supercomplex) and induces their dissociation from the core, although in the absence of PSI

they remain in the grana; (iii) trimer M and S get phosphorylated but they do not dissociate

from the core. Here we show that both the LHCII of the PSII supercomplex as well as the

“extra” trimers get phosphorylated upon stat 1 to state 2 transition but that phosphorylated

trimer M and S remain associated with the PSII core. And this in WT plants, when the

docking side on PSI is available. This is in agreement with previous results (Larsson, Sundby

et al. 1987) which showed that there are at least two pools of LHCII that can be

phosphorylated, of which only one moves to the stroma lamellae. Our data clearly indicate

that the association of trimer-S and M with the PSII core is too strong to be broken by

phosphorylation in the membrane. However, phosphorylation can be instrumental in

breaking the weak interactions between the PSII supercomplex and the “extra” LHCII

trimers, thus allowing the latter to move to PSI.

4.3 Remodeling of PSII supercomplexes and megacomplexes is not required for state

transitions

In this work we have investigated several important aspects of state transition in higher

plants at the level of supercomplexes, looking at their composition and organization.

However, we would like to point out here that other mechanisms, such as thylakoid

membrane reorganization, might also occur during state transitions. The results of this study

are summarized in Fig. 5: the composition and organization of the PSII supercomplexes does

not show significant differences between state 1 and state 2. The LHCII that migrates to the

stroma lamellae in state 2 is a subpopulation of “extra” LHCII which is probably located close

to the grana margin in state 1, as suggested by the absence of changes in the PSII density in

the grana in the two conditions.

Page 135: University of Groningen From Photosystem I to Photosystem

134

Figure 5. Schematic model of LHCII phosphorylation during state transitions in higher plants. The PSI and PSII core are depicted in yellow, in PSII trimers M and S in blue, the “extra” LHCII trimers in green, the minor PSII antenna in orange and the Lhca PSI antenna in red. Phosphorylation of LHCII is indicated with a P. The “extra” LHCIIs are mainly associated with the PSII C2S2M2 supercomplexes located nearby the stroma. Phosphorylation of PSII supercomplex LHCII trimers and “extra” LHCII trimers (St2) results in the migration of the latter to PSI in the stroma, while the C2S2M2 supercomplex remains intact. The mobile LHCII trimer will most likely associate with PSI complexes located nearby the grana.

In state 2, not only are the “extra” mobile LHCII phosphorylated, but also the LHCII trimers

which are associated with the core to form the PSII supercomplexes. This indicates that

phosphorylation is not sufficient to induce the movement of LHCII from grana to stroma as

previously suggested (Lunde, Jensen et al. 2000). Moreover, the phosphorylation is not

sufficient to detach the LHCII trimers from the PSII supercomplex. The population of “extra”

LHCII is weakly connected with PSII supercomplexes as indicated by the fact that they

transfer energy relatively slowly to the reaction center of PSII in state 1 (Wientjes, van

Amerongen et al. 2013), and that it has thus far not been possible to purify a PSII

supercomplex from plants with these trimers attached (Caffarri, Kouril et al. 2009). It is

therefore likely that this connection is sufficiently weak to be broken by phosphorylation.

At the same time it can be expected that phosphorylated LHCII has a higher affinity for PSI,

such that the effect is synergetic. The presence of the migrating LHCII in the grana margins

also means that the LHCII do not have to migrate very far during state transitions, thus

limiting the possible danger caused by the presence of “free” pigment-protein complexes in

the membrane. In addition, we have recently shown that in basically all natural light

conditions for plants, a part of the mobile LHCII population is associated with PSI (Wientjes,

Page 136: University of Groningen From Photosystem I to Photosystem

135

van Amerongen et al. 2013). It can thus be concluded that the association of LHCII with PSI

has evolved from a short-term response as observed in green algae, where it involves up to

80% of the LHCII population (Delosme, Olive et al. 1996) and has a large effect on the PSII

organization (Iwai, Takahashi et al. 2008), to a long-term response in plants, where LHCII is

an antenna of PSI in most light conditions.

Acknowledgment

We thank Laura M. Roy for help with the membrane preparations and for helpful comments

on the manuscript. This work was supported by the Netherlands Organisation for Scientific

Research (NWO), Earth and Life Sciences (ALW), through a Vici grant and the ERC

starting/consolidator grant 281341 (to R.C.) and the grant No. ED0007/01/01 Centre of the

Region Haná for Biotechnological and Agricultural Research (to R.K.).

Page 137: University of Groningen From Photosystem I to Photosystem

136

References

Allen, J. F. (1992). "How Does Protein-Phosphorylation Regulate Photosynthesis." Trends Biochem Sci 17(1): 12-17.

Allen, J. F. (1992). "Protein-Phosphorylation in Regulation of Photosynthesis." BBA 1098(3): 275-335. Allen, J. F., J. Bennett, et al. (1981). "Chloroplast Protein-Phosphorylation Couples Plastoquinone

Redox State to Distribution of Excitation-Energy between Photosystems." Nature 291(5810): 25-29.

Allen, J. F. and J. Forsberg (2001). "Molecular recognition in thylakoid structure and function." Trends Plant Sci 6(7): 317-326.

Allen, J. F. and T. Pfannschmidt (2000). "Balancing the two photosystems: photosynthetic electron transfer governs transcription of reaction centre genes in chloroplasts." Philosophical Transactions of the Royal Society of London Series B-Biological Sciences 355(1402): 1351-1357.

Allen, J. F., S. Santabarbara, et al. (2011). "Discrete Redox Signaling Pathways Regulate Photosynthetic Light-Harvesting and Chloroplast Gene Transcription." Plos One 6(10).

Bassi, R. and P. Dainese (1992). "A Supramolecular Light-Harvesting Complex from Chloroplast Photosystem-Ii Membranes." Eur J Biochem 204(1): 317-326.

Bassi, R., G. Giacometti, et al. (1988). "Changes in the organization of stroma membranes induced by in vivo state 1-state 2 transition." BBA 935: 152-165.

Bellafiore, S., F. Bameche, et al. (2005). "State transitions and light adaptation require chloroplast thylakoid protein kinase STN7." Nature 433(7028): 892-895.

Bennett, J. (1991). "Protein-Phosphorylation in Green Plant Chloroplasts." Annu Rev Plant Phys and Plant Mol Biol 42: 281-311.

Bonaventaru, C. and J. Myers (1969). "Fluorescence and Oxygen Evolution from Chlorella Pyrenoidosa." BBA 189(3): 366-383.

Broess, K., G. Trinkunas, et al. (2008). "Determination of the Excitation Migration Time in Photosystem II - Consequences for the Membrane Organization and Charge Separation Parameters." BBA-Bioenergetics 1777(5): 404-409.

Caffarri, S., R. Kouril, et al. (2009). "Functional Architecture of Higher Plant Photosystem II Supercomplexes." Embo J 28(19): 3052-3063.

Chow, W. S., A. Melis, et al. (1990). "Adjustments of photosystem stoichiometry in chloroplasts improve the quantum efficiency of photosynthesis." Proc Natl Acad Sci U S A 87(19): 7502-7506.

Croce, R., g. Canino, et al. (2002). "Chromophore organization in the higher-plant photosystem II antenna protein CP26." Biochemistry 41(23): 7334-7343.

Croce, R., G. Zucchelli, et al. (1996). "Excited state equilibration in the Photosystem I light-harvesting I complex: P700 is almost isoenergetic with its antenna." Biochemistry 35(26): 8572-8579.

Damkjaer, J. T., S. Kereiche, et al. (2009). "The Photosystem II Light-Harvesting Protein Lhcb3 Affects the Macrostructure of Photosystem II and the Rate of State Transitions in Arabidopsis." Plant Cell 21(10): 3245-3256.

Dekker, J. P. and E. J. Boekema (2005). "Supramolecular Organization of Thylakoid Membrane Proteins in Green Plants." BBA-Bioenergetics 1706(1-2): 12-39.

Delosme, R., J. Olive, et al. (1996). "Changes in light energy distribution upon state transitions: An in vivo photoacoustic study of the wild type and photosynthesis mutants from Chlamydomonas reinhardtii." Biochimica Et Biophysica Acta-Bioenergetics 1273(2): 150-158.

Dietzel, L., K. Braautigam, et al. (2011). "Photosystem II Supercomplex Remodeling Serves as an Entry Mechanism for State Transitions in Arabidopsis." Plant Cell 23(8): 2964-2977.

Galka, P., S. Santabarbara, et al. (2012). "Functional Analyses of the Plant Photosystem I-Light-Harvesting Complex II Supercomplex Reveal That Light-Harvesting Complex II Loosely Bound to Photosystem II Is a Very Efficient Antenna for Photosystem I in State II." Plant Cell 24(7): 2963-2978.

Grieco, M., M. Tikkanen, et al. (2012). "Steady-State Phosphorylation of Light-Harvesting Complex II Proteins Preserves Photosystem I under Fluctuating White Light." Plant Physiology 160(4): 1896-1910.

Page 138: University of Groningen From Photosystem I to Photosystem

137

Hankamer, B., J. Nield, et al. (1997). "Isolation and biochemical characterisation of monomeric and dimeric photosystem II complexes from spinach and their relevance to the organisation of photosystem II in vivo." Eur J Biochem 243(1-2): 422-429.

Hogewoning, S. W., E. Wientjes, et al. (2012). "Photosynthetic quantum yield dynamics: from photosystems to leaves." Plant Cell 24(5): 1921-1935.

Iwai, M., Y. Takahashi, et al. (2008). "Molecular remodeling of photosystem II during state transitions in Chlamydomonas reinhardtii." Plant Cell 20(8): 2177-2189.

Jansson, S. (1999). "A guide to the Lhc genes and their relatives in Arabidopsis." Trends Plant Sci 4(6): 236-240.

Jarvi, S., M. Suorsa, et al. (2011). "Optimized Native Gel Systems for Separation of Thylakoid Protein Complexes: Novel Super- and Mega-Complexes." Biochem J 439: 207-214.

Kouril, R., G. T. Oostergetel, et al. (2011). "Fine structure of granal thylakoid membrane organization using cryo electron tomography." Biochimica Et Biophysica Acta-Bioenergetics 1807(3): 368-374.

Kouril, R., E. Wientjes, et al. (2013). "High-light vs. low-light: Effect of light acclimation on photosystem II composition and organization in Arabidopsis thaliana." Biochim Biophys Acta 1827(3): 411-419.

Kouril, R., A. Zygadlo, et al. (2005). "Structural characterization of a complex of photosystem I and light-harvesting complex II of Arabidopsis thaliana." Biochemistry 44(33): 10935-10940.

Kovacs, L., J. Damkjaer, et al. (2006). "Lack of the light-harvesting complex CP24 affects the structure and function of the grana membranes of higher plant chloroplasts." Plant Cell 18(11): 3106-3120.

Kyle, D. J., L. A. Staehelin, et al. (1983). "Lateral mobility of the light-harvesting complex in chloroplast membranes controls excitation energy distribution in higher plants." Archives of Biochemistry and Biophysics 222(2): 527-541.

Larsson, U. K., B. Jergil, et al. (1983). "Changes in the lateral distribution of the light-harvesting chlorophyll-a/b--protein complex induced by its phosphorylation." Eur J Biochem 136(1): 25-29.

Larsson, U. K., C. Sundby, et al. (1987). "Characterization of 2 Different Subpopulations of Spinach Light-Harvesting Chlorophyll a/B-Protein Complex (Lhc-Ii) - Polypeptide Composition, Phosphorylation Pattern and Association with Photosystem-Ii." BBA 894(1): 59-68.

Lunde, C., P. E. Jensen, et al. (2000). "The PSI-H subunit of photosystem I is essential for state transitions in plant photosynthesis." Nature 408(6812): 613-615.

Melis, A. (1991). "Dynamics of Photosynthetic Membrane-Composition and Function." BBA 1058(2): 87-106.

Minagawa, J. (2011). "State transitions-The molecular remodeling of photosynthetic supercomplexes that controls energy flow in the chloroplast." BBA-Bioenergetics 1807(8): 897-905.

Murata, N. (1969). "Control of Excitation Transfer in Photosynthesis .I. Light-Induced Change of Chlorophyll a Fluorescence in Porphyridium Cruentum." BBA 172(2): 242-&.

Nelson, N. and C. F. Yocum (2006). "Structure and Function of Photosystems I and II." Annu Rev Plant Biol 57: 521-565.

Oostergetel, G. T., W. Keegstra, et al. (1998). "Automation of specimen selection and data acquisition for protein electron crystallography." Ultramicroscopy 74(1-2): 47-59.

Peter, G. F. and J. P. Thornber (1991). "Biochemical-Composition and Organization of Higher-Plant Photosystem-Ii Light-Harvesting Pigment-Proteins." J Biol Chem 266(25): 16745-16754.

Pribil, M., P. Pesaresi, et al. (2010). "Role of Plastid Protein Phosphatase TAP38 in LHCII Dephosphorylation and Thylakoid Electron Flow." Plos Biology 8(1).

Schagger, H. (2006). "Tricine-SDS-PAGE." Nature Prot. 1(1): 16-22. Shapiguzov, A., B. Ingelsson, et al. (2010). "The PPH1 phosphatase is specifically involved in LHCII

dephosphorylation and state transitions in Arabidopsis." Proc Natl Acad Sci U S A 107(10): 4782-4787.

Tikkanen, M., M. Nurmi, et al. (2008). "Phosphorylation-dependent regulation of excitation energy distribution between the two photosystems in higher plants." BBA-Bioenergetics 1777(5): 425-432.

Wientjes, E., H. van Amerongen, et al. (2013). "LHCII is an Antenna of Both Photosystems After Long-term Acclimation " BBA-Bioenergetics 1827: 420-426.

Page 139: University of Groningen From Photosystem I to Photosystem

138

Wientjes, E., H. van Amerongen, et al. (2013). "Quantum Yield of Charge Separation in Photosystem II: Functional Effect of Changes in the Antenna Size upon Light Acclimation." J. Phys. Chem. B DOI: 10.1021/jp401663w.

Page 140: University of Groningen From Photosystem I to Photosystem

139

Summary

Photosynthesis is arguably the most important biological process on Earth. Directly

or indirectly it provides reduced carbon required for the survival of virtually all life on our

planet and supplies molecular oxygen. The evolution of photosynthesis started early in

Earth’s history, over 3.8 billion years ago. The first photosynthetic organisms established the

composition of the biosphere that led to the development of advanced life forms.

In oxygenic photosynthetic organisms - cyanobacteria, algae and higher plants - light

reactions of photosynthesis are catalyzed by four highly specialized membrane complexes:

Photosystem I, Photosystem II, cytochrome b6f and ATP synthase. In these processes, light

energy is absorbed and converted into chemical energy - ATP and NADPH, which are used

for the fixation of carbon dioxide into organic matter. As a side product, oxygen is produced.

The light reactions of photosynthesis have been extensively studied over the decades. Large

part of the research has made use of model systems, like the higher plant Arabidopsis

thaliana and the unicellular green alga Chlamydomonas reinhardtii. Despite the fact that

studies on these organisms have strongly contributed to the understanding the

photosynthetic process, many open questions remain regarding the molecular mechanisms

of light harvesting and its regulation. This is particularly true for C. reinhardtii for which also

a biochemical characterization of the photosynthetic complexes is missing: what is the

composition of its photosynthetic complexes? How do the organization of Photosystem I

and Photosystem II change in different conditions? What is their antenna size? Also

processes, such as state transitions, photoprotection and photoinhibition are not fully

understood at the molecular level yet.

The aim of this thesis was to answer these questions, providing comprehensive structural

and functional information about the photosynthetic apparatus of C. reinhardtii,

in particular on the composition and organization of Photosystem I and Photosystem II

(chapter 2 and 3) and on their flexibility in different environmental conditions (chapter 4

and 5). A combination of biochemical, proteomic, and spectroscopic techniques, as well as

electron microscopy was used. The information obtained represents the basis for a full

understanding of light-harvesting and its regulation in this alga. Novel PSI and PSII

complexes have been purified and characterized, allowing us to relate the protein

composition and organization to the functional behaviour of the supercomplexes.

Page 141: University of Groningen From Photosystem I to Photosystem

140

Moreover, the comparison of the photosystems of green algae and higher plants gives

insight regarding adaptation and evolution of the photosynthetic apparatus.

The understanding of the molecular basis of the photosynthetic process and the control of

light harvesting under different environmental conditions may be used to improve the

quality of human life. The “know-how” of Chlamydomonas photosynthesis can be used to

increase photosynthetic efficiency, for biofuels production and can also guide the design of

new and efficient ways to collect and use solar energy.

Chapter 2 describes the study of the composition of Photosystem I of C. reinhardtii and the

supramolecular organization of its outer antenna (Lhcas). Using a two steps purification

procedure, homogeneous PSI-LHCI complexes have been isolated and the main properties

of these PSI-LHCI supercomplexes have been revealed. The data show that the largest

purified complex contains a core complex composed of two large subunits – PsaA and PsaB,

surrounded by several smaller subunits, and a light harvesting antenna system composed of

nine Lhca gene products. A projection map at 15 Å resolution obtained by electron

microscopy shows that the nine Lhcas are organized on one side of the core in a double half-

ring arrangement. A series of stable disassembled intermediates of PSI-LHCI was also

purified. The analysis of these complexes suggests the sequence of assembly/disassembly of

PSI and shows that PSI-LHCI of C. reinhardtii is less stable than the complex of higher plants.

This can explain the large variation in the antenna composition of PSI-LHCI of C. reinhardtii

found in the literature. The analysis of subcomplexes with reduced antenna size allows the

determination of the position of Lhca2 and Lhca9 and leads to a proposal for a model of the

organization of the Lhcas within the PSI-LHCI supercomplex. Lhca2 and Lhca9 (the red-most

antenna complexes), although present in the largest complex in 1:1 ratio with the core, are

only loosely associated with it. Lhca4, Lhca5 and Lhca6 are located in the outer half-ring of

the supercomplex, while Lhca1, Lhca3, Lhca7 and/or Lhca8 compose the inner half-ring.

In chapter 3 the focus is on the organization of C.r.PSII supercomplexes and its association

with light-harvesting antenna (LHCII). Using very mild solubilisation several PSII

supercomplexes with different antenna size have been purified. Electron microscopy

analysis of single particles revealed PSII complexes organization ranging from dimeric core

(C2), C2S2 complex to a novel PSII-LHCII supercomplex which contains three LHCII trimers per

monomeric core (C2S2M2N2). A projection map at 13 Å resolution was obtained which

Page 142: University of Groningen From Photosystem I to Photosystem

141

allowed determining the position and the orientation of the Lhcb’s within the

supercomplex. The position and orientation of the S trimer are the same as in plants; trimer

M is rotated by 45° degrees and the additional trimer (named here LHCII-N), which is taking

the position occupied in plants by CP24, is directly associated with the core. The analysis of

supercomplexes with different antenna size suggests that LhcbM5 is part of the “extra”

LHCII pool not directly associated with the supercomplex. It is also shown that

Chlamydomonas LHCII has a slightly lower chlorophyll a/b ratio than the complex from

plants and a blue shifted absorption spectrum. Finally the data indicate that there are

at least six LHCII trimers per dimeric core in the thylakoid membranes.

Taken together the results of chapters 2 and 3 show that the light-harvesting capacity of

both Photosystem I and Photosystem II in C. reinhardtii is larger than that of the

photosystems of plants, but the similar enlargement for both photosystems allows them to

maintain the absorption balance and thus linear electron transfer. However, fluctuation of

light quality and quantity leads to an imbalance in the energy distribution between the two

photosystems. To adapt to changing light conditions, the light harvesting capacity of PSI and

PSII is modulated, to maximize photosynthesis and minimize photodamage. State transitions

represent a key regulatory process of the light reactions in C. reinhardtii. In this process

energy distribution between photosystems is balanced by the reversible association of

mobile LHCII with either PSII (state 1) or PSI (state 2). In C.r. it is believed that 80% of LHCII

undergoes state transitions. However, the consequences of state transitions on the

structural organization and the functionality of the photosystems of C. reinhardtii have not

been fully elucidated yet.

Successful purification of C.r. photosystems during state transitions (chapter 4) allows

performing structural and functional studies on these complexes. A projection map of PSI

complexes in state 2 at 20 Å resolution reveals the structural organization of the complex.

It is shown that PSI in state 2 is able to bind two LHCII trimers and one monomer, most likely

CP29, in addition to its nine Lhcas. This is the largest PSI complex ever observed, far larger

than that of higher plants where only one LHCII trimer was found to be associated with PSI

in state 2. Moreover, all PSI-bound Lhcs are efficient in transferring energy to PSI. The data

also show that PSII-LHCII can be purified from cells in state 2, although in a smaller amount

than in state 1, indicating that part of the supercomplexes are intact in those conditions.

Page 143: University of Groningen From Photosystem I to Photosystem

142

This suggests that only part of the Lhcb population can participate in state transitions. The

number is far lower than the 80% reported in the literature. It is also shown that contrary to

all expectations, several Lhcb subunits can associate with PSI in their unphosphorylated

form, and to PSII in their phosphorylated form.

As in C. reinhardtii also in higher plants, the transition from state 1 to state 2 is triggered by

phosphorylation of LHCII. Although several models have been proposed of how state

transitions influence PSII organization, the identity of the trimer that moves from PSII to PSI

is still under debate. Chapter 5 focuses on the reversible association of LHCII with

Photosystems I and II during state transitions in higher plants. Two LHCII trimers (called

trimers M and S) are part of the PSII supercomplex, while up to two more are loosely

associated with PSII in state 1 in higher plants (called “extra” trimers). Here we show that

the LHCII from the “extra” pool migrates to PSI in state 2, thus leaving the PSII supercomplex

and the semi-crystalline PSII arrays intact. In state 2 not only is the mobile LHCII

phosphorylated, but also the LHCII in the PSII supercomplexes. This demonstrates that PSII

phosphorylation is not sufficient for disconnecting LHCII trimers S and M from PSII and for

their migration to PSI.

The comparison of the results presented in chapter 4 and 5 indicates that in C. reinhardtii

most of the Lhcb complexes are able to act as antenna for both PSI and PSII, while in A.

thaliana the mobile pool of LHCII consists of the “extra” (non-PSII supercomplex) LHCII. The

results also indicate that the well-accepted models in which phosphorylation leads to

dissociation of Lhcb complexes from PSII and to their association with PSI, model needs to

be revised.

Page 144: University of Groningen From Photosystem I to Photosystem

143

Samenvatting

Fotosynthese is waarschijnlijk het belangrijkste biologische proces op aarde. Direct

of indirect verschaft het gereduceerde koolstof die onmisbaar is voor het voortbestaan van

vrijwel al het leven op onze planeet. Het voorziet bovendien de zuurstof consumerende

organismen van moleculaire zuurstof. De evolutie van fotosynthese begon vroeg in de

geschiedenis van de aarde, meer dan 3,8 miljard jaar geleden. De eerste fotosynthetische

organismen hebben de samenstelling van de biosfeer gevestigd wat tot de ontwikkeling van

complexe levensvormen heeft geleid.

In zuurstof-fotosynthetische organismen – cyanobacteriën, algen en vaatplanten, worden de

lichtreacties van fotosynthese door vier hooggespecialiseerde membraancomplexen

gekatalyseerd: Fotosysteem I, Fotosysteem II, cytochroom b6f eb ATP synthase. In dit proces

wordt het licht geabsorbeerd en omgezet in chemische energie – ATP en NADPH, die

gebruikt worden voor de fixatie van koolstofdioxide naar de organische materie. Als

bijproduct wordt zuurstof geproduceerd.

De lichtreacties van fotosynthese zijn door de jaren heen uitgebreid onderzocht. Het

merendeel van het onderzoek heeft gebruikgemaakt van modelorganismen, zoals de

vaatplant Arabidopsis thaliana en de eencellige groenwier Chlamydomonas reinhardtii. De

studie van deze organismen heeft bijgedragen tot een beter begrip van hoe de processen

van fotosynthese verlopen. Er bestaan echter nog steeds onbeantwoorde vragen over de

fotosynthese van C. reinhardtii, zoals: Wat is de samenstelling van de fotosynthetische

complexen? Hoe verandert de organisatie van Fotosysteem I (PSI) en Fotosysteem II (PSII) in

verschillende omstandigheden? Wat is hun antennemaat? Ook processen als

toestandsovergangen, fotobescherming of fotoinhibitie vergen gedetailleerder onderzoek.

De antwoorden op bovenstaande vragen werden behandeld in dit proefschrift. We hebben

ons gericht op de studie van de fotosynthetische complexen in Chlamydomonas reinhardtii.

Het doel was om uitgebreide structurele en functionele informatie over Fotosysteem I en

Fotosysteem II te krijgen (hoofdstuk 2 en 3) en om hun flexibiliteit in verschillende

omgevingsomstandigheden te testen (hoofdstuk 4 en 5). Om de meest gedetailleerde

informatie te krijgen hebben we gebruikgemaakt van zowel een combinatie van

biochemische, proteomische en spectroscopische technieken als elektronenmicroscopie. De

verkregen informatie vormt de basis voor een volledig begrip van light-harvesting en de

regulatie daarvan in deze alge. De vergelijking met de gegevens uit vaatplanten helpt de

evolutie van de light-harvesting regulatie te begrijpen. Nieuwe PSI en PSII complexen zijn

Page 145: University of Groningen From Photosystem I to Photosystem

144

gezuiverd en gekarakteriseerd, waardoor het mogelijk was de proteïnesamenstelling en -

organisatie aan het functioneel gedrag van de supercomplexen te relateren.

Het begrip van de moleculaire grondslagen van het fotosynthetische proces en de controle

van light-harvesting onder verschillende omgevingsomstandigheden kan gebruikt worden

om de kwaliteit van het mensenleven te verbeteren. De ‘know-how’ van de fotosynthese

van Chlamydomonas kan gebruikt worden om de fotosynthetische efficiency, cultuurgroei

en het rendement te vergroten. Als gevolg daarvan kan de productie van biobrandstoffen

worden verhoogd. Bovendien kan het helpen nieuwe, efficiënte manieren van verzamelen

en gebruiken van zonne-energie te ontwikkelen.

Hoofdstuk 2 beschrijft het onderzoek naar de samenstelling van Fotosysteem I van

C. reinhardtii en de supramoleculaire organisatie van zijn buitenste antenne (Lhcas). Door

middel van een tweestappen zuiveringsprocedure zijn er homogene PSI-LHCI complexen

geïsoleerd. De hoofdeigenschappen van dit PSI - LHCI supercomplex zijn onthuld. De

gegevens laten zien dat het grootste gezuiverde complex een kerncomplex bevat bestaand

uit twee grote subeenheden – PsaA en PsaB, omringd door enkele kleinere subeenheden, en

uit een light-harvesting antenne bestaande uit negen Lhca-genproducten. Een beeld met de

resolutie van 15 Å verkregen met behulp van elektronenmicroscopie laat zien dat negen

Lhca’s aan één kant van de kern in een dubbele halfringconfiguratie zijn georganiseerd. Ook

een reeks van stabiele sub-complexen van PSI-LHCI is gezuiverd. De analyse van deze

complexen suggereert de assemblage/disassemblage sequentie van PSI en laat zien dat PSI-

LHCI van C. reinhardtii minder stabiel is dan het vaatplantencomplex. Dat kan de in de

literatuur aangetroffen grote variatie in de antenne-samenstelling van PSI-LHCI van C.

reinhardtii verklaren. De analyse van subcomplexen met gereduceerde antennemaat maakt

het mogelijk om de positie van Lhca2 en Lhca9 te bepalen en leidt tot een voorstel van een

organisatiemodel van de Lhca’s binnen het PSI-LHCI supercomplex. Lhca2 en Lhca9, hoewel

aanwezig in het grootste complex in 1:1 verhouding met de kern, houden daar slechts een

los verband mee. Lhca4, Lhca5 en Lhca 6 bevinden zich in de buitenste halfring van het

supercomplex, terwijl Lhca1, Lhca3, Lhca7 en/of Lhca8 in de binnenste halfring.

Hoofdstuk 3 concentreert zich op de organisatie van C.r. PSII supercomplexen en de relatie

hiervan met de light-harvesting antenne (LHCII). Met behulp van zeer milde solubilisatie zijn

enkele PSII supercomplexen met verschillende antennemaat gezuiverd. Een

elektronenmicroscopische analyse van een enkelvoudig deeltje onthulde organisatie van

PSII complexen, uiteenlopend van dimerische kern (C2), C2S2 complex, tot een nieuw PSII-

LHCII supercomplex dat drie LHCII trimeren per monomerische kern bevat (C2S2M2N2). Er is

een beeld met de resolutie van 13 Å verkregen waardoor het mogelijk was de positie en

Page 146: University of Groningen From Photosystem I to Photosystem

145

oriëntatie van de Lhcb’s binnen het supercomplex te bepalen. De positie en oriëntatie van

de S-trimeer zijn dezelfde als in planten; trimeer M is geroteerd over 45 graden en de

additionele trimeer (hier LHCII-N genoemd), die de positie inneemt die in planten door CP24

wordt ingenomen, houdt een direct verband met de kern. De analyse van supercomplexen

met verschillende antennemaat suggereert dat LhcbM5 een deel is van de ‘extra’ LHCII

groep die geen direct verband houdt met het supercomplex. Tevens zien we dat

Chlamydomonas LHCII heeft een wat lagere chlorofyl a/b ratio dan de complexen van

planten en een blauw-verschoven absorptiespectrum. Ten slotte wijzen de gegevens uit dat

er zich ten minste zes LHCII trimeren per dimerische kern in de thylakoïdemembranen

bevinden.

Al de resultaten van hoofdstukken 2 en 3 samen laten zien dat de light-harvesting capaciteit

van beide Fotosysteem I en Fotosysteem II in C. reinhardtii groter is dan die van de

fotosystemen van planten. De vergelijkbare uitbreiding voor beide fotosystemen stelt ze

echter in staat om de absorptie-evenwicht te onderhouden en dus ook lineaire

elektronenoverdracht. Schommeling van lichtkwaliteit en -kwantiteit leidt echter tot een

onevenwichtigheid in de energiedistributie tussen de twee fotosystemen. Om aan de

veranderende lichtomstandigheden aan te passen, is de light-harvesting capaciteit van PSI

en PSII gemoduleerd om fotosynthese te maximaliseren en fotobeschadiging te

minimaliseren. Toestandsovergangen vormen een belangrijk regelgevend proces van de

lichtreacties in C. reinhardtii. In dit proces wordt de energiedistributie tussen fotosystemen

in evenwicht gehouden door de omkeerbare verbinding van mobiele LHCII met of PSII

(toestand I), of PSI (toestand II). Er wordt aangenomen dat in C.r. 80% van LHCII

toestandsovergang ondergaat. De gevolgen van toestandsovergangen voor de structurele

organisatie en de functionaliteit van de fotosystemen van C. reinhardtii zijn echter nog niet

volledig opgehelderd.

Geslaagde zuivering van C.r. fotosystemen tijdens toestandsovergangen (hoofdstuk 4)

maakt structureel en functioneel onderzoek van deze complexen mogelijk. Een beeld van

PSI complexen in toestand 2 met de resolutie van 20 Å onthult de structurele organisatie

van het complex. We zien dat PSI in toestand 2 de mogelijkheid heeft om twee LHCII

trimeren en een monomeer, hoogstwaarschijnlijk CP29, te binden als aanvulling op zijn

negen Lhca’s. Dat is het grootste PSI complex dat ooit is geobserveerd, veel groter dan dat

van vaatplanten waar slechts één LHCII trimeer een verband met PSI bleek te houden in

toestand 2. Bovendien zijn alle aan het PSI gekoppelde Lhcs’ efficiënt in energie-overdracht

naar PSI. De gegevens laten ook zien dat PSII-LHCII gezuiverd kan worden van cellen in

toestand 2, hoewel in een kleiner aantal dan in toestand 1, wat uitwijst dat een deel van de

Page 147: University of Groningen From Photosystem I to Photosystem

146

supercomplexen in die omstandigheden intact is. Dit suggereert dat maar een deel van de

populatie van Lhcb deel kan nemen aan toestandsovergangen. Dit aantal is veel lager dan de

80% die in de literatuur wordt vermeld. We zien bovendien dat, in tegenstelling tot alle

verwachtingen, enkele Lhcb subeenheden een verband kunnen houden met PSI in hun

ongefosforyleerde vorm, en met PSII in hun gefosforyleerde vorm.

Zoals in C. reinhardtii, ook in vaatplanten wordt de overgang van toestand 1 naar 2 gestart

door fosforylering van LHCII. Hoewel er enkele modellen zijn voorgesteld van hoe de

organisatie van PSII door toestandsovergangen wordt beïnvloed, blijft het karakter van de

trimeer, die van PSII naar PSI overgaat nog altijd onderwerp van discussie. Hoofdstuk 5

concentreert zich op de omkeerbare verbinding van LHCII met Fotosystemen I en II tijdens

toestandsovergangen in vaatplanten. Twee LHCII trimeren (genoemd trimeer M en S)

maken deel uit van het PSII supercomplex, terwijl tot twee meer trimeren een los verband

houden met PSII in toestand 1 in vaatplanten (genoemd ‘extra’ trimeren). We laten hier zien

dat de LHCII uit de ‘extra’ groep zich in toestand 2 naar PSI verplaatst en dus zowel PSII

supercomplex als de semi-kristallijn PSII reeksen intact behoudt. In toestand 2 wordt niet

alleen de mobiele LHCII gefosforyleerd maar ook de LHCII in de PSII supercomplexen. Dit

laat zien dat PSII fosforylering niet genoeg is om LHCII trimeren S en M van PSII te

ontkoppelen en om de migratie daarvan naar PSI mogelijk te maken.

De vergelijking van de in hoofdstukken 4 en 5 gepresenteerde resultaten wijst uit dat in C.

reinhardtii de meeste Lhcb complexen als antenne voor beide PSI en PSII blijken te kunnen

fungeren, terwijl in A. thaliana de mobiele groep van LHCII uit de ‘extra’ (niet-PSII

supercomplex) LHCII bestaat. De resultaten wijzen ook uit dat de algemeen erkende

modellen, waarin fosforylering tot de dissociatie van Lhcb complexen uit PSII en tot

associatie daarvan met PSI leidt, moeten worden herzien.

Page 148: University of Groningen From Photosystem I to Photosystem

147

Acknowlegments

My PhD life was an exciting jouney. During this time, all my successes and failures in

the lab, as well as in daily life thought me a lot and made me a better scientist and a better

person. For me this journey is coming to an end. It would not have been possible to get here

without the support of several people. I would like to express my sincere gratitude to all of

them.

First of all I would like to thank Prof. Roberta Croce for giving me the opportunity to join her

group. You were the key figure throughout my PhD research. I could not have imagined

having a better supervisor and mentor for my PhD study. Your door was always open for

me. I really appreciate your patience, enthusiasm and immense knowledge. Without your

guidance this thesis would not have been possible.

I would like to express my special gratitude and thanks to the people from the Electron

Microscopy group at the University of Groningen who made an enormous contribution to

this thesis. I gratefully thank Prof. Egbert J. Boekema for helpful discussions and for reading

my works during all these years. I thank Mariam, Sathish and Roman. Without you no one

could see these beautiful complexes.

I am deeply grateful to Francesca and Emilie for helping me at the beginning of my research.

You shared your knowledge teaching me how to work with photosynthetic complexes.

I have been extremely fortunate to have worked with such talented scientists like you.

I express my gratitude to my collaborators: Dr. Fabrizia Fusetti and Alicja for the mass

spectrometry experiments. Your work and comments have been of a great help.

I appreciate the time spending on reading this thesis and the feedback offered by the

members of the reading committee: Prof. E. J. Boekema, Prof. B. Robert and Prof. S. J.

Marrink.

I thank Laura Roy for proofreading the introduction and for her helpful comments on the

manuscript.

I thank Marija Smits for the permission to use her beautiful painting as a cover for this book.

Page 149: University of Groningen From Photosystem I to Photosystem

148

Thanks to Hilda and Regine for the administrative assistance, for the help with forms, rules,

laws and all related paperwork.

I thank my fellow labmates of the Photosynthesis group at VU Amsterdam for the good

atmosphere. A special thanks to Clo for stimulating discussions. Thanks to Nikki, Iryna,

Ludwig, Alberto, Lucyna, Michael for all the fun we have had in and outside the lab in the

last years.

I would like to thank all my friends. First I would like to thank Agata and Mariusz for taking

care of me during all the years I spent in Groningen. We had rally good time. The fun and

support you have provided were priceless.

To the “polish mafia”: Justyna, Marysia, Hania, Gosia, Jacek, Wojtek, Piotrek, Slawek,

Maciek thank you. For our Sunday dinners and crazy nights we spent together. It has been a

pleasure to meet all of you.

Kochani rodzice. Wasza miłość i wyrozumiałość pozwoliła mi iść przez życie pełne przeszkód,

ale także przez życie obfitujące w chwile szczęścia i radości. Wasza pomoc w zdobywaniu

wiedzy i doświadczenia życiowego jest nieoceniona.

A special thanks to my sister Kate, who has always supported, encouraged and believed in

me.

Finally, I dedicate the last few lines to the dearest person to me, my beloved Justyna.

Without you this effort would have been worth nothing. Thank you for your continued love

and support.