Development and dynamics of the photosynthetic apparatus in purple phototrophic bacteria

Development and dynamics of the photosynthetic apparatus in purple phototrophic bacteria

    Development and dynamics of the photosynthetic apparatus in purple phototrophic bacteria Robert A. Niederman PII: DOI: Reference: S0...

1MB Sizes 0 Downloads 75 Views

    Development and dynamics of the photosynthetic apparatus in purple phototrophic bacteria Robert A. Niederman PII: DOI: Reference:

S0005-2728(15)00221-2 doi: 10.1016/j.bbabio.2015.10.014 BBABIO 47549

To appear in:

BBA - Bioenergetics

Received date: Revised date: Accepted date:

22 August 2015 22 October 2015 25 October 2015

Please cite this article as: Robert A. Niederman, Development and dynamics of the photosynthetic apparatus in purple phototrophic bacteria, BBA - Bioenergetics (2015), doi: 10.1016/j.bbabio.2015.10.014

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

PT

ACCEPTED MANUSCRIPT

RI

Development and Dynamics of the Photosynthetic Apparatus in Purple

NU

SC

Phototrophic Bacteria

MA

Robert A. Niederman*

D

Department of Molecular Biology and Biochemistry, Rutgers, The State University of New

AC CE P

TE

Jersey, 604 Allison Road, Piscataway, New Jersey 08854-8082, United States

For Special Issue of BBA Bioenergetics devoted to: ―Organization and dynamics of bioenergetic systems in bacteria‖

.

*E-mail address: [email protected]

Keywords: Atomic force microscopy, Bacterial photosynthesis; light-harvesting; reaction center; membrane assembly; membrane dynamics

1

ACCEPTED MANUSCRIPT Highlights Proteomics showed assembly factor enrichment at membrane growth initiation sites.



CCCP triggered an accumulation of factors functioning early in the assembly process.



A cooperative assembly mechanism for the RC-LH1 complex was demonstrated by AFM.



Fluorescence relaxation showed a slowing of RC electron transfer as levels of LH2 increased.



This is thought to arise mainly from an increased distance for cytochrome c2 diffusion.

SC

RI

PT



NU

Abbreviations: AFM, atomic force microscopy; BChl, bacteriochlorophyll a; -OG, n-octyl β-D-

MA

glucopyranoside; CCCP, carbonyl-cyanide m-chlorophenyl-hydrazone; CM, cytoplasmic membrane; EM, electron micrograph; DOC, deoxycholate; FM, maximal fluorescence; FV,

D

variable fluorescence; FV /FM, the quantum yield of the primary charge separation; ICM,

TE

intracytoplasmic membrane; LH, light harvesting; LH1, core light-harvesting complex; LH2, peripheral light-harvesting complex; p, light-harvesting bacteriochlorophyll connectivity, RC,

AC CE P

photochemical reaction center; functional absorption cross-section; ET, the rate of RC electron transfer turnover.

2

ACCEPTED MANUSCRIPT ABSTRACT The purple bacterium Rhodobacter sphaeroides provides a useful model system for studies of the

PT

assembly and dynamics of bacterial photosynthetic membranes. For the nascent developing

RI

membrane, proteomic analyses showed an ~2-fold enrichment in general membrane assembly factors, compared to chromatophores. When the protonophore carbonyl-cyanide m-

SC

chlorophenyl-hydrazone (CCCP) was added to an ICM inducing culture, an ~2-fold elevation in

NU

spectral counts vs. the control was seen for the SecA translocation ATPase, the preprotein translocase SecY, SecD and SecF insertion components, and chaperonins DnaJ and DnaK, which

MA

act early in the assembly process. It is suggested that these factors accumulated with their nascent polypeptides, as putative assembly intermediates in a functionally arrested state. Since in

TE

D

Synechocystis PCC 6803, a link has been established between Chl delivery involving the highlight HilD protein and the SecY/YidC-requiring cotratranslational insertion of nascent

AC CE P

polypeptides, such a connection between BChl biosynthesis and insertion and folding of nascent Rba. sphaeroides BChl binding proteins is likely to also occur. AFM imaging studies of the formation of the reaction center (RC)-light harvesting 1 (LH1) complex suggested a cooperative assembly mechanism in which, following the association between the RC template and the initial LH1 unit, addition of successive LH1 units to the RC drives the assembly process to completion. Alterations in membrane dynamics as the developing membrane becomes filled with LH2-rings was assessed by fluorescence induction/relaxation kinetics, which showed a slowing in RC electron transfer rate thought to mainly reflect alterations in donor side electron transfer. This was attributed to an increased distance for electron flow in cytochrome c2 between the RC and cytochrome bc1 complexes, as suggested in current structural models.

3

ACCEPTED MANUSCRIPT 1. Introduction……………………………………………………………………………………5 2. Proteomic analyses of membrane development in Rhodobacter sphaeroides……………...6

PT

2.1 Analysis of changes in membrane organization during a shift from high to low light intensity

RI

and induction of ICM formation at low oxygen tension……………….………………………........…7 2.2. Identification of early factors involved in the assembly of proteins of the bacterial

SC

photosynthetic apparatus…………………………………………………………………………………...8

NU

3. Role of assembly factors in targeting and binding of BChl(Chl) to apoproteins followed by their oligomerization and organization into supramolecular structures……………..15

MA

3.1 YidC is involved in binding of Chl to assembling apoproteins in cyanobacteria……………. 15 3.2 Roles for complex-specific assembly factors in the formation of bacteriochlorophyll a-protein

TE

D

complexes………………………………………………………………………………………...16 3.3 Oligomerization of LH1 and assembly of the RC-LH1 core complex as observed in AFM

AC CE P

studies…………………………………………………………………………………………………. 20 3.4. LH2 and RC-LH1play roles in establishing ICM architecture………………...……….…..22 4. Dynamics of developing ICM as examined by fluorescence emission analyses………….25 4.1 Bacteriochlorophyll a-based fluorescence induction/relaxation kinetics……………..…..…25 4.2 Alterations in membrane dynamics as assessed by fluorescence induction/relaxation analyses……………………………………………………………………………………………..……....29 4.3. Searching for a mechanism explaining the decreased reaction center electron transfer turnover rate during greening and adaptation to low light intensity……………………...…….31 References…………………………………………………………………………………...…..36

4

ACCEPTED MANUSCRIPT 1. Introduction Following a period of explosive growth of research on the structure and function of the

PT

intracytoplasmic photosynthetic membrane (ICM) of anoxygenic purple phototrophic bacteria

RI

[1], largely stemming from the determination of the first atomic resolution crystal structures of the photochemical reaction center (RC) [2-5] and light-harvesting (LH) antenna [6-8] proteins, a

SC

new research era has emerged in which the in situ supramolecular organization and dynamics of

NU

photosynthetic unit components within the membrane are now undergoing detailed investigation. This epoch began just over a decade ago with the advent of atomic force microscopy (AFM)

MA

providing native surface views of the ICM, obtained at submolecular resolution (see 9-12 for reviews). In addition to providing the earliest in situ topological views of any multi-component

TE

D

energy transducing membrane [13], AFM has revealed a wide diversity of species-dependent topological arrangements of closely packed antenna and RC structures, capable of fulfilling the

AC CE P

basic requirements for efficient collection, transmission and trapping of radiant energy. In ICM patches from Rhodobacter sphaeroides grown at low light intensity [14], a highly organized architecture was observed consisting of well-ordered, interconnected rows of dimeric RC-LH1 core complexes intercalated with areas of peripheral LH2 antenna, coexisting with extensive LH2-only domains (Fig. 1A,B). In most other species studied such as Rhodospirillum photometricum, a much less regular organization was seen [15], with mixed regions of LH2 antenna and RC-LH1 core structures intermingled with large, paracrystalline LH2 domains (Fig. 1D-F). AFM together with fluorescence induction/relaxation kinetic analyses have provided the structural and functional paradigms for functional proteomic approaches to ICM development in Rba. sphaeroides in response to alterations in light intensity and oxygen tension [20-28]. The

5

ACCEPTED MANUSCRIPT fluorescence yield properties of cells adapting to both lowered light intensity and low oxygen levels that induce ICM development, have revealed a direct relationship between the slowed rate

PT

of RC electron transfer turnover and the increasing LH2/LH1 molar ratios and the functional

RI

absorption cross-section [23,24,27-29]. This has been ascribed to the imposition of constraints upon the free diffusion of redox species between the cytochrome bc1 complex and the RC as the

SC

ICM bilayer is expanded and becomes densely packed with LH2 rings. This possibility was

NU

confirmed by a comparison of ICM patches from cells grown at high and low light intensity by high-resolution AFM [14], in which the increased LH2 levels in the low-light ICM were shown

MA

to form the densely packed LH2-only domains that were absent in the high-light cells (Fig.1). These LH2-only domains represent the light-responsive antenna complement formed under low

TE

D

illumination. These structural and functional results have been largely confirmed by the proteomic analyses described below.

AC CE P

2. Proteomic analyses of membrane development in Rhodobacter sphaeroides Rba. sphaeroides has numerous advantages that provide an ideal model system for examining the development and assembly of photosynthetic membranes. This metabolically versatile purple chlorophototrophic bacterium has an accessible genetic system together with an ICM that provides an unparalleled variety of biochemical, spectroscopic and ultrastructural probes. When grown photoheterotrophically (anaerobically in the light), the levels of LH2 relative to the RC-LH1 core complexes are related inversely to light intensity [30] and changes in illumination levels have offered a valuable paradigm for examining the differential biosynthesis of photosynthetic complexes. While under chemoheterotrophic growth conditions, ICM formation is repressed, lowering the oxygen partial pressure initiates gratuitous induction of ICM formation in the dark (greening) by invagination of the CM [31]. Under these circumstances, sequential

6

ACCEPTED MANUSCRIPT assembly of the LH and RC complexes occurs in which the functionally essential RC-LH1 complex is synthesized first, followed by the peripheral LH2 antenna [29].

PT

2.1 Analysis of changes in membrane organization during a shift from high to low light intensity

RI

and induction of ICM formation at low oxygen tension

By availing ourselves of these capabilities, a proteomic analysis of the ICM development

SC

process was performed in which the temporal expression of membrane proteins was followed

NU

both during transitions from high to low light intensity [21,23,25,27] and the induction of ICM formation at low aeration [32]. Membrane protein components, which have assessable

MA

ultrastructural and functional correlates, are spatially localized in distinct membrane regions arising both from areas of nascent photosynthetic membrane development as well as the mature

TE

D

ICM, isolated in the respective upper pigmented and ICM vesicles (chromatophore) bands after rate-zone sedimentation of cell-free extracts in sucrose density gradients. This facilitates the

AC CE P

assessment of the proteomes arising from distinct membrane regions at various stages of ICM development. For proteomic analysis, the chromatophore and nascent membrane fractions were subjected to clear native electrophoresis for isolation of bands containing the LH2 and RC–LH1 core complexes. In chromatophores, the proteomic profile of the trypsin-digested LH2 gel bands (Fig. 2), revealed increasing levels of LH2 spectral counts relative to those of the RC–LH1 complex as ICM development proceeded during a light-intensity downshift [21], essentially confirming the results of the spectral and AFM analyses (see below). A large array of other associated proteins was observed in the LH2 band, including high spectral counts for the F1FOATP synthase subunits and the cytochrome bc1 and pyridine nucleotide transhydrogenase complexes, as well as the hypothetical protein RSP6124. Significant levels of RSP6124 were previously reported [20] and although topological predictions suggested a lack of transmembrane

7

ACCEPTED MANUSCRIPT domains, its spectral counts were especially elevated in the LH2 band, nearly equaling those of the LH2 polypeptides during the first 3 h following the shift from high to low light intensity

PT

(Fig.2). Consequently, it is possible that RSP6124 may play a transient role in the assembly or

RI

function of the LH2 antenna.

In contrast to chromatophores, the RC-LH1 band from the purified upper-pigmented fraction

SC

was enriched in cytoplasmic membrane (CM) markers (Fig. 3) including electron transfer

NU

proteins (2-fold) and transport proteins (3-fold), as well as general membrane protein assembly factors (2-fold)[32]. The enriched assembly factors included preprotein translocases SecY, YidC

MA

and YajC, together with the bacterial type 1 signal peptidase and twin arg translocation subunit TatA. These data confirm the origin of the upper pigmented band from the peripheral respiratory

TE

D

membrane and sites of active CM invagination that give rise to the ICM, in which preferential assembly of the RC-LH1 complex occurs. The complete upper pigmented fraction isolated from

AC CE P

cells undergoing semiaerobic induction of ICM formation for 12 h [32] showed a much larger enrichment in CM markers (~4 fold in electron transfer components, ~14-fold in transporters), consistent with the markedly decreased levels of RC-LH1 and LH2 proteins. This fractions also showed an ~3-fold increase in general membrane protein assembly factors and detectable levels of LH complex specific assembly factors. Substantial levels of GroEL heat shock chaperonin were also found, however, reflecting the limited cell division occurring in the concentrated cell preparations used in this induction procedure. Under such circumstances, GroEL promotes the folding of nascent proteins to counteract stress-induced protein denaturation and aggregation. 2.2. Identification of early factors involved in the assembly of proteins of the bacterial photosynthetic apparatus Early factors driving the assembly of ICM proteins were identified by a proteomic approach

8

ACCEPTED MANUSCRIPT in which the lipid-soluble protonophore carbonyl-cyanide m-chlorophenylhydrazone (CCCP) was utilized to block the insertion of membrane proteins during the induction of ICM formation

PT

under low aeration in concentrated Rba. sphaeroides cell suspensions. CCCP acts as an

RI

uncoupler of electron transfer from ATP synthesis [36] by inhibiting the formation of the electrochemical proton gradient generated during electron transport by the respiratory chain

SC

carriers [37]. Integration of CCCP into the membrane bilayer in an anionic form results in proton

NU

acquisition from the periplasmic space. After diffusion of the protonated CCCP form back across the membrane, protons are discharged into the cytoplasm, collapsing the electrochemical proton

MA

gradient and leaving CCCP in the anionic state. Multiple repetitions of this process serve to short-circuit the respiratory chain.

D

In the closely-related purple bacterium, Rba. capsulatus, concentrations of CCCP blocking

polypeptide subunit into the ICM for LH2 complex assembly [38]. It

AC CE P

integration of the LH2-

TE

cytochrome c2 export to the periplasmic space (50 M) were also shown to abolish the

was concluded that maintenance of an electrochemical proton gradient is required for translocation of the nascent pigment-binding proteins. This represents step two of the dynamic sequence of steps delineated [39] in guiding of nascent pigment binding proteins across the membrane and assembling them into functional photosynthetic complexes: (i) cotranslational chaperonin assisted folding of the nascent polypeptides into membrane insertion-competent conformations and binding to the cytoplasmic surface of the membrane; (ii) electrochemical proton gradient driven insertion and integration into the membrane; (iii) delivery and binding of the BChl and carotenoid chromophores; (iv) oligomerization of the individual protomers into oligomeric annular structures; (v) supramolecular organization into functional photosynthetic units. Here, we will concentrate largely on the second step and bring some new perspectives to

9

ACCEPTED MANUSCRIPT steps three and four. With regard to the general membrane assembly factors, studies on the formation of the LH1

PT

complex in Rba. capsulatus using a membrane coupled cell-free translation system, demonstrated that depletion of chaperonin DnaK significantly decreased the rate of synthesis and

RI

membrane insertion of both the LH1-α and –β polypeptides [40,41]. Removal of the GroEL

SC

chaperonin affected stable LH1 insertion into the coupled membranes, and expectedly, the

NU

assembly process was also dependent on the presence of unidentified membrane-bound factors. The following procedures were used to assess if the CCCP-induced collapse of the

MA

electrochemical proton gradient resulted in the accumulation of assembly factors, in association with nascent polypeptides as predicted by step 2 of the assembly pathway [39]. ICM formation

TE

D

was gratuitously induced by shifting BChl-depleted, concentrated suspensions of aerobically grown cells to semiaerobic conditions. Over the first 8 h, a vast increase in the number of both

AC CE P

the RC-LH1 core and peripheral LH2 antenna complexes occurred, where the preferential appearance of the core particles is followed by an accelerated accumulation of LH2 [29,32,42], as reflected by a >4–fold increase in the molar LH2/LH1 ratio over this time period. Since limited cell division occurred in the concentrated cell suspensions, functional changes largely reflected accelerated synthesis of pigments and their binding proteins in existing cells. Therefore, this system was ideal for examining the effects of inhibition of the electrochemical proton gradient on the active assembly of pigment-protein complexes. After the addition of CCCP at 8 h, fluorescence induction/relaxation measurements showed a marked reduction in the quantum yield of the primary charge separation, as assessed from the Fv/Fm values (see Fig. 7 legend for definitions). A >4-fold slowing of the rate of RC electron transfer turnover (2.4 to 11.5 ms) occurred within the first hour after addition of the protonophore. Since multiple cycles of CCCP

10

ACCEPTED MANUSCRIPT inhibition lead to a collapse of the proton gradient and short circuit the respiratory chain, the resulting short-circuit would be expected to affect proton release phases in the Q-cycle of the bc1

PT

complex, limiting donor side electron transfer to the RC [43]. The physiological parameters

RI

assessed by these fluorescence emission measurements confirmed that 50 M CCCP should be

SC

more than sufficient to block the membrane-potential requirement for insertion of nascent polypeptide chains into the membrane bilayer.

NU

To test the effects of CCCP on protein insertion into the ICM, chromatophore fractions isolated from the inducing cells at 4 h following the onset of CCCP treatment, and from the

MA

untreated 12-h control cells were subjected to clear native electrophoresis for LC-MS/MS analysis of trypsinized gel bands [32]. The effects of CCCP treatment on the LH, RC and related

TE

D

energy transduction complexes from major gel bands are shown in Fig. 4. CCCP markedly blocked the membrane insertion of polypeptides of the LH2, RC-LH1 and F1FO-ATPase

AC CE P

complexes, especially as sampled in the respective gel bands in which they were highly enriched. A 2.7-fold reduction in spectral counts was observed for the LH2 complex from the bottom (LH2) band, consistent with the accelerated rate of LH2 synthesis and assembly seen at this stage of the induction process. In contrast, 1.7–1.9-fold reductions were found for the RC-LH1 complex in the top (RC-LH1) band, and in the upper intermediate band for the F1-ATP and FO-ATP synthases. Likewise, the pyridine nucleotide transhydrogenase, an additional energy transducing ICM-associated protein complex, showed a spectral count reduction of 1.7-fold, averaged over all-four gel bands. Accordingly, this complex, along with the RC-LH1 core and F1FO-ATPase complexes, with spectral count decreases in the of 1.7-1.9-fold range, reflect blockage by CCCP of the membrane insertion of nascent polypeptides of complexes with rates of assembly at a basal level during this stage of ICM induction, rather than at the accelerated rate

11

ACCEPTED MANUSCRIPT seen for the LH2 complex. Moreover, these results provide novel in vivo evidence that confirms the necessity of the maintenance of an electrochemical proton gradient as an essential element in

PT

membrane translocation of the integral LH2 apoproteins [38] and extends these findings to the

RI

RC-LH1, transhydrogenase and F1FO-ATPase complexes. A CCCP-induced reduction in spectral counts was also observed for enzymes of BChl and carotenoid synthesis (1.6- and 1.7-fold,

SC

respectively). Although the majority of these enzymes represent soluble proteins, they may be

NU

degraded in the absence of appropriate associations with membrane components required for coordinating their respective pathways of pigment biosynthesis with concomitant formation of

MA

pigment binding apoproteins of the LH and RC complexes. It still seemed possible that the decline in spectral counts in the energy transducing

TE

D

membrane proteins reflected a slowing of protein synthesis arising from decreases in ATP levels caused by the effects of electrochemical proton gradient dissipation on oxidative

AC CE P

phosphorylation. This was shown not to be the case by the similar levels of spectral counts between the CCCP-treated and -untreated chromatophores for 35 selected soluble proteins, representing enzymes of carbohydrate and amino acid metabolism (Fig. 4). This indicated that the rate of protein biosynthesis was unaffected during the 4-h protonophore treatment. Therefore, the CCCP-induced major decline in levels of the energy-transducing membrane protein complexes can be attributed to protonophore-induced blockage in the electrochemical protein gradient necessary for their membrane insertion rather than a general arrest in protein biosynthesis. As expected, CCCP treatment resulted in the accumulation of general membrane assembly factors in the unfractionated upper-pigmented fraction, as shown in the proteomic profile seen in Fig. 5. An ~2-fold elevation in the spectral counts vs. control levels was observed for the SecA

12

ACCEPTED MANUSCRIPT translocation ATPase, the preprotein translocase SecY, SecD and SecF insertion components, and chaperonins DnaJ and DnaK. Since each of these components functions in early stages of the

PT

membrane insertion process, they may have accumulated in association with nascent

RI

polypeptides in a functionally arrested state. The resulting transient assembly intermediates, formed as a consequence of CCCP-induced blockage of the electrochemical gradient, apparently

SC

exist in a state that is not readily dissociated or degraded. In contrast, YidC and YagC, which

NU

play roles in later stages of the Sec translocase dependent prokaryotic membrane assembly process [44], showed a CCCP-induced decline in their levels and appear to have undergone

MA

degradation. A varied level of elevation of CCCP-induced spectral counts was observed for the highly conserved SPFH domain protein [45], the FtsH metalloproteinase/chaperonin [46], and

TE

D

HflC, a member of the SPFH group which functions along with HflK as a modulator of FtsH activity. FtsH has been shown to cross-link with YidC and HflK/C, suggesting that these proteins

AC CE P

are linked to play a role in quality control of the membrane protein integration process [47]. Other components involved in early aspects of membrane protein assembly that were slightly elevated by CCCP treatment included the signal recognition particle homologue Ffh/SRP54 [48] and signal peptidase l [49]. Assembly factors that failed to accumulate, and showed essentially no changes in levels in the upper-pigmented fraction after CCCP treatment, included complex specific assembly factors (not shown) functioning in cofactor binding and assembly steps, following the initial insertion of nascent polypeptides into the membrane. Among these components with the highest spectral counts were the cytochrome c-type biogenesis protein, cycH [50] iron-sulfur cluster-binding protein [51] and LH1 assembly factors [52]. Overall, these data suggest that nascent polypeptides destined for BChl binding and with roles in energy

13

ACCEPTED MANUSCRIPT transducing complexes, are inserted cotranslationally into the membrane in Sec-dependent manner.

PT

The observation in AFM topographs of ―captive‖ proteins in association with LH1 rings

RI

formed in an LH-only Rba. sphaeroides strain (Fig. 6)[53], provides in vivo structural support for the existence of membrane protein-assembly factor intermediates as suggested from the

SC

proteomic analysis of CCCP-treated cells [32]. The presence of complete LH1 rings still

NU

associated with their assembly factors is thought to reflect the absence of an RC template, leading to LH1 formation around other proteins of appropriate dimensions. Integral membrane

MA

proteins of sufficient size to serve in this capacity include BChl synthase (BchG), catalyzing bacteriochlorophyllide esterification and LhaA, a RC-LH1 core assembly factor with a possible

TE

D

role in delivery of BChl to apoproteins (see below). Not only have proteomic approaches applied to the upper-pigmented fraction facilitated

AC CE P

elucidation of relative levels of associated general membrane assembly factors and their LH complex-specific counterparts [32], but also it was possible to assess the dynamics of the ICM assembly process in which general assembly factors involved in early assembly stages accumulated as putative transient intermediates as a result of CCCP-induced blockage of the membrane insertion of nascent polypeptides. Distinct transient assembly intermediates have been shown to accumulate upon disruption of photosystem assembly in the cyanobacterium Synechocystis PCC 6803 [54], leading to the identification of Psb28, a factor involved in Chl biosynthesis as well as formation of apoprotein subunits of Chl-binding proteins (the PSII antenna complex CP47, PSI RC core proteins PsaA and PsaB)[55]. Subsequently, the multifunctional assembly factor Psb27 was found to associate with the unassembled PSII antenna complex CP43, larger CP43 containing complexes, as well as the PsaB [56]. The associations

14

ACCEPTED MANUSCRIPT forming these transient assembly intermediates were demonstrated through comigration in native gel electrophoresis, copurification with His-tagged PSI and a yeast two-hybrid assay.

PT

3. Role of assembly factors in targeting BChl(Chl) to apoproteins followed by their

RI

oligomerization and organization into supramolecular structures

3.1 YidC is involved in binding of Chl to assembling apoproteins in cyanobacteria

SC

In an investigation of mechanisms involved in the delivery of Chl to nascent photosystem

NU

apoproteins, a strain of Synechocystis PCC 6803 was constructed for pull down assays to detect proteins associated with Chl synthase (ChlG), the terminal enzyme of Chl biosynthesis [57,58].

MA

Among the associated proteins was HliD, a tightly bound protein induced at high light intensity, also associated with the putative PSII assembly factor Ycf39 and the YidC insertase. In addition,

TE

D

YidC was shown by immunoblotting to be associated with ribosomal subunits. Deletion of hilD resulted in elevated levels of the Chl G substrate, chlorophyllide, suggesting a role for HilD in

AC CE P

the formation of ChlG complexes containing Ycf39 and other assembly factors. These studies established a link between Chl delivery by HilD and the SecY/YidC-requiring insertion of nascent photosystem polypeptides into cyanobacterial photosynthetic membranes [57]. Evidence for such a possible BChl-protein assembly supercomplex is discussed below. The polypeptide foldase activity of Yid C may be a crucial player in both insertion of Chl into nascent pigment binding proteins and establishment of suitable Chl binding sites. The identification of YidC together with ribosomal subunits as part of the ChlG)/pigment apoprotein assembly supercomplex is important, since ribosomal pausing has been shown to occur at distinct sites during elongation of the D1 subunit of PSII [59] and may be coordinated with Chl binding. It was suggested [57] that Yid C arranges the growing nascent polypeptide chains into a series of programmed configurations amenable to Chl insertion and that the concerted pausing by

15

ACCEPTED MANUSCRIPT the ribosome provides the stops along the growing nascent polypeptide chain for correct attachment of Chl molecules, as stipulated by the adjacent ChlG, ultimately forming the correctly

PT

folded and occupied binding pockets of active Chl containing proteins.

RI

Among the earliest functions established for YidC in Escherichia coli is its role as a membrane chaperonin, facilitating Sec translocase-dependent and -independent lateral bilayer

SC

movement of nascent polypeptides [60]. In the case of the SecYEG translocon-dependent

NU

pathway, YidC occupies the lateral gate of SecYEG and becomes displaced in a sequential manner by nascent polypeptides, thereby facilitating the lipid bilayer transfer of transmembrane

MA

segments [61]. YidC also functions in the folding of nascent polypeptides and in promoting helix-helix interactions in conjunction with the SecYEG machinery [62]. The Sec translocase-

TE

D

independent YidC pathway drives the membrane integration of the subunits a and c of the FO sector of the F1FO ATPase [63] and Subunit II (CyoA) of the cytochrome o complex [64], a

AC CE P

conserved function in respiratory chain biogenesis. YidC binds directly to the SecDFYajC in forming a heterotetrameric accessory complex [65] linking the Sec translocase complex to YidC, while SecD and SecF also participates in SecY independent insertion of proteins into the bilayer. Moreover, a holo-translocon consisting of SecYEG-SecDF-YajC-YidC along with an electrochemical proton gradient drives both the co-translational membrane protein insertion and posttranslational secretion processes [66]. Thus, YidC acts as a multifunctional assembly factor and is capable of catalyzing a multitude of activities to assist in the targeting and binding of cofactors during the insertion of nascent apoproteins into the membrane in a fully folded and functionally active state. 3.2 Roles for complex-specific assembly factors in the formation of bacteriochlorophyll a-protein complexes

16

ACCEPTED MANUSCRIPT A supercomplex of protein assembly factors can also be envisioned for the implementation of a concerted mechanism for BChl synthase (BchG)/BChl delivery/apoprotein assembly, thus

PT

minimizing free pigment exposure by rapid transfer between the synthase and a nearby

RI

translocon channel, thus avoiding the threat of photooxidative damage to the developing photosynthetic apparatus. Among the complex-specific assembly factors encoded by genes

SC

located in the photosynthetic gene cluster of Rba. sphaeroides are LhaA, an LH1-specific

NU

assembly factor, and the related LH2-specific factor PucC [52]. These proteins are members the major facilitator superfamily of transport proteins, forming part of the putative ―chlorophyll

MA

delivery branch.‖ A role for LhaA in BChl cofactor assembly was suggested from immunoblotting analyses, indicating an enrichment of both LhaA and the BChl synthase protein

D

in the Rba. sphaeroides upper-pigmented fraction [67,68] that was shown to contain nascent

TE

photosynthetic membranes that serve as a ―hotspot‖ for both BChl [67] and pigment binding

AC CE P

protein biosynthesis [See 27,28 for reviews]. It is essential that the targeting process deliver the nascent apoproteins to the membrane at the precise location and time of BChl availability through coordination of protein integration with pigment synthesis. Thus, a supercomplex can be envisioned that consists in part of BChl synthase (BChG) and such putative BChl delivery proteins as LhaA and PucC, together with the translocase component YidC. This would establish the appropriate protein architecture for pigment binding and lead to fully folded and assembled photosynthetic complexes. Since apoproteins of the RC-LH1 and LH2 complexes were the major components undergoing synthesis in the oxygen-depleted, concentrated cell suspensions during CCCP treatment [32], it is possible that the Sec translocase participates in their membrane insertion, insofar as putative early assembly intermediate were found to accumulate in the absence of an electrochemical proton gradient. These apparent intermediates contained the SecA

17

ACCEPTED MANUSCRIPT translocation ATPase, the preprotein translocase SecY, and the SecD and SecF insertion components.

PT

Genes for additional factors with photosynthetic complex assembly roles [52] have been

RI

identified within the photosynthesis gene clusters of Rba sphaeroides and Rba. capsulatus. Included are the genes encoding PufQ, a putative BChl carrier protein [69-71], PuhB, an

SC

apparent dimeric RC assembly factor which secondarily affects LH1 assembly [72], PuhC,

NU

required for optimal RC and LH1 levels, and thought to affect RC/LH1/PufX reorganization [73], and PuhE, a negative modulator of BChl synthesis [74]. In Rba sphaeroides, transcripts of

MA

pufQ, puhB and puhE were detected in a DNA microarray analysis of the oxygen and lightregulation of transcriptome expression [75]. Evidence of a role for PufQ in BChl

TE

D

biosynthesis/delivery comes from genetic studies in Rba. capsulatus, which showed a linear relation between pufQ expression levels and the formation of BChl, as well as conserved BChl

AC CE P

binding sequences [69]. PufQ was identified in Rba. capsulatus membranes by immunoblotting, where detectable levels were correlated with BChl synthesis after a transfer from aerobic to semiaerobic growth conditions [70]. Moreover, PufQ was also shown to have an affinity for the BChl precursor protochlorophyllide, when reconstituted into liposomes [71]. Further studies are necessary before a definitive role can be established for any of these putative assembly factors. It is important to note that the pufQ gene is at the 5’ end of the PufQBALMX operon encoding PufQ, the LH1- and -polypeptides, respectively, the RC-L and RC-M subunits and PufX. The PufX protein promotes the dimerization of the RC-LH1 core complex [76,77], in which it is interspersed within the LH1 ring a 1:1 stoichiometry with the RC [77]. The dimerization process prevents complete encirclement of LH1 around the RC, forming an LH1-based portal for efficient ubiquinone/ubiquinol exchange between the reaction center QB site and the cytochrome

18

ACCEPTED MANUSCRIPT bc1 complex [78]. PufX is also implicated in the supramolecular organization of the Rba. sphaeroides ICM through alleged promotion of a uniform long-range order to the RC-LH1

PT

complex, as suggested by linear dichroism (LD) spectroscopy [79,80]. LD interrogates the

RI

difference in absorption between horizontally and vertically polarized light, relative to the

SC

orientation of membrane preparations that are ordered macroscopically in polyacrylamide gels. This facilitates determination of the transition-dipole moment orientations of relevant

NU

chromophores, viz. the accessory RC BChl, the BChl special pair and LH1 and LH2 BChls. In both LD and dark-minus-light difference LD spectra of oriented membranes containing the

MA

native RC-LH1-PufX complex, the RC was distinguished by distinctive accessory BChl and BChl special pair spectral features, not found with RC-LH1 complexes lacking PufX [79].

D

-

TE

As first seen in EMs of freeze fracture replicas of elongated tubular membranes in the LH2

Rba. sphaeroides strain M21, dimeric RC–LH1 complexes form highly ordered arrays that are

-

AC CE P

arranged helically along the longitudinal axes of the tubules [82]. LD measurements of oriented LH2 membranes showed that the RC BChls are locked in a unique, uniform orientation within individual RCs over the entire intact membrane. [79]. Oriented wild-type chromatophore preparations also exhibited a remarkable degree of structural orientation in which all RC-LH1 core structures were shown to align in regular arrays with all RC BChl transition moments oriented in the same direction, despite the presence of LH2 as the predominant antenna complex in these native membranes. Moreover, the relative contributions of the LH1 and LH2 complexes to the overall LD spectrum in chromatophore vesicles in which the LD of LH2 was reduced relative to that in the LH1 absorption peak, showed that LH2 preferentially resides in highly curved parts of ICM membrane vesicles, since the LD and absorption peaks were equalized in flattened membrane patches [80].

19

ACCEPTED MANUSCRIPT While considerable information on how the supramolecular structure and organization of photosynthetic unit cores is established has arisen from these studies, it has recently been

PT

suggested from variations in the level of supramolecular organization of the Rba. sphaeroides

RI

ICM induced by alterations in PufX, that PufX may play an indirect role in the ordering of RC-LH1 dimers over the membrane [83]. Instead long-range order may stem from the role of

SC

PufX in promoting an appropriate dimer configuration with the creation of ordered RC–LH1

NU

domains depending not only on the presence of RC–LH1 dimers in their appropriate bent conformation, but also on the structure of LH1 on the outside of the complex. Thus, the sole

MA

influence that PufX exerts on membrane ordering may be via the facilitation of the dimerization process in which the dimer has one long axis of symmetry and is bent along the dimer interface

TE

D

[84].

3.3 Oligomerization of LH1 and assembly of the RC-LH1 core complex as observed in AFM

AC CE P

studies

In an AFM imaging study of the assembly of single RC-LH1 molecules using membranes from an LH1 only mutant [53], a series of aberrant forms of the LH1 complex in which the basic repeating unit (LH1-11(BChl)2) was seen to assume ring structures of variable size, elliptical structures, spirals and arcs that appear to represent assembly products (Fig. 6). The spiral complexes occur as a result of a failure in the closing of the ring, while the short arcs, first observed in lithium dodecyl sulfate/polyacrylamide gel electrophoresis [82], arise from premature termination of the LH1 complex assembly process. Occasionally, captive proteins were observed that were enclosed within the LH1 ring (Fig. 6) and as noted above, they could correspond on the basis of their size to assembly intermediates in which BChl synthase or complex specific assembly factors such as LhaA are bound and still associated. The presence of

20

ACCEPTED MANUSCRIPT such intermediates may reflect LH1 formation around proteins other than the RC template, which is absent in this strain [53]. A mixture of fully formed RC-LH1 complexes, emptied LH1 rings

PT

and isolated RCs was imaged (Fig. 6C) when formation of LH1 was restricted by a lowering of the cognate pufBA transcripts levels through removal of the stem-loop structure between the pufA

SC

RI

and pufL genes and progressive truncation at the C-terminus of the LH1- polypeptide. Overall, the resulting assortment of structures suggested a cooperative mechanism for assembly of the

NU

RC-LH1 complex in which once an association occurs between the RC and the first LH1 repeating unit, RC associations with successive LH1 subunits drive the LH1 ring assembly

MA

process to completion. Partial LH1 rings were not observed in the LH1-limited constructs (Fig. 6C), establishing that the necessity of an RC template for LH1 ring assembly. These AFM

TE

D

studies essentially confirm previous results (85) from an examination of ICM induction under low aeration in which the RC-H subunit was detected first, followed by PufX and thereafter by a

AC CE P

low level of LH1- and -subunits, which successively increased. The assembly process was proposed to reflect the initial formation of a RC-(LH1-11(Bchl)2)-PufX complex, with progressive addition of LH1-11(BChl)2 subunits, that eventually encircle the RC. Forces guiding the assembly of the LH2 complex, which unlike LH1 consists of an ―empty‖ ring structure, have been studied within the native ICM of Rsp. photometricum by AFM imaging of single molecules, combined with force measurements [86]. The binding of LH2 protomers was found to be fairly weak, indicating the importance of LH2 ring architecture in stabilizing the structure of the LH2 complex. In addition, intermolecular forces were shown to influence LH2 structural and functional integrity, as demonstrated from AFM unfolding kinetics in which the G for unfolding of an LH2 subunit consisting of an LH2-11(BChl)3 structure was estimated to be 222 kJ mol-1. Since a similar G might be anticipated for unfolding of an LH1-11(BChl)2

21

ACCEPTED MANUSCRIPT subunit, the oligomerization of an LH1-only complex would be expected to be driven by small cumulative inter-subunit forces that facilitate ring formation (53). In the case of a spiral complex,

PT

crowding pressure from adjacent rings might be expected to misalign the nascent complex so that

RI

additional subunits are assembled to form a spiral. Overall, the studies on forces driving the assembly of the LH2 complex [86] showed that: (i) weakly bound subunits are capable of self-

SC

assembling to form rings; (ii) the stability of the complex is guarantied by the ring-shape alone;

NU

(iii) the individual subunits are essentially solid and their stability further depends on the intramembrane molecular environment.

MA

3.4. LH2 and RC-LH1 play roles in establishing ICM architecture The discovery of a vesicular ICM similar to that of the wild-type in the Rba. sphaeoroides

TE

D

NF57 LH2-only mutant (i.e. lacking the RC and LH1 complexes) suggested that the LH2 oligomeric ring structure possesses the membrane bending properties responsible for the ICM

AC CE P

vesiculation process [87,88]. On the other hand, a tubular ICM system was found in the M21 mutant lacking LH2 that was sometimes seen running between two cells appearing to be arrested in cell division (not shown). The observed ultrastructural differences suggested that in the absence of LH2, ICM morphogenesis is incomplete and is arrested at a tubular stage. The presence of vesicular ICM in the NF57 strain indicated that LH2 alone is both necessary and sufficient for vesicular ICM formation and that membrane morphogenesis can be completed in the absence of RC-LH1 core particles. As noted in section 3.2, the RC-LH1 complex exists in Rba. sphaeroides mainly as an Sshaped dimeric structure. The dimerization process is facilitated by the PufX protein, which serves as an integral component of the core complex (RC2LH1- pair28PufX2)[77]. In an effort to determine the extent to which the dimeric RC-LH1 complex controls the morphology of the

22

ACCEPTED MANUSCRIPT ICM, a 3-D RC-LH1-PufX model was constructed from single particle EM analysis of negatively stained complexes [84]. The two halves of the dimer were inclined toward one

PT

another on the periplasmic aspect of the complex, creating an uncommon V-shaped structure

RI

measuring ~200 A° along the long dimer axis and 100 A° along the short axis. The strikingly bent modeled structure has revealed the basis for the imposition of membrane curvature by the

SC

core complex and the ability to drive the formation of budded, tubular membrane structure,

NU

together with the arrangement of dimeric RC-LH1 complexes in highly ordered helical arrays along the tubules and the ordered packing of the dimers in 2D crystals. It can also be envisioned

MA

that RC-LH1 dimers utilize their distinctive curvature properties to facilitate their aggregation into the rows of RC-LH1 complexes observed uniquely in AFM topographs of Rba. sphaeroides

TE

D

[13] to facilitate excitation energy transfer from LH1 to the RC. In an investigation of the membrane bending effects of the highly bent RC-LH1 complex,

AC CE P

molecular dynamics simulations were employed for constructing an all-atom structural model within the EM density map [89]. A membrane of high local curvature was simulated which agreed well with the size of RC-LH1 tubules and further demonstrated how the intrinsic geometric shape of RC-LH1 dimers, consisting of a structure bent at the dimer interface, propagates the extended organization of a tubular ICM. In a similar approach, using the atomic resolution X-ray structure of the Rhodoblastus acidophilus LH2 and a homology-modeled Rba. sphaeroides LH2 structure [90], membrane embedded arrays of LH2 rings structures were found to assume a dynamic curvature, consistent with the putative LH2 membrane bending properties. An array of seven LH2 rings, as observed in AFM for LH2 complexes of Rsp. photometricum membranes (Fig. 1E), gave simulations in which these structures were seen to tilt away from neighboring complexes on the cytoplasmic side which created a curvature for the LH2

23

ACCEPTED MANUSCRIPT aggregates, that consequently curving the membrane. Thus, the aggregation dependent properties of individual LH2 rings were proposed to account for the membrane curving properties.

PT

Although a collections of seven rings grouped together is often encountered in Rsp.

RI

photometricum membranes, the ICM of this organism consists of regular bundles of flattened thylakoid-like discs rather than vesicular ICM. Moreover, while EMs of high-light grown Rba.

SC

sphaeroides in which LH2 concentration is low, clearly show a vesicular ICM [14], the LH2

NU

rings therein do not not form large aggregates, but instead are present mainly as individual structures in dimeric association or short single-file rows (Fig. 1C, 14). Also, simulations based

MA

on the atomic resolution LH2 structure of Rbl. acidolphilus showed the same arrangement of seven grouped LH2 rings as for Rba. sphaeroides LH2 and no substantial differences in

TE

D

membrane curvature were simulated, despite the existence of a lamellar ICM in Rbl. acidolphilus. Thus, while LH2 is clearly a membrane-bending protein, the overall mechanism of

AC CE P

membrane bending and vesiculation remain elusive. In this connection, the possibility of a role for protein-lipid interactions in establishing ICM architecture in Rba. sphaeroides has has been suggested from the finding of a putative LH2-lipid binding site and enrichment of the non-bilayer phospholipid phosphatidylethanolamine in the lipids associated at the boundary with LH2 [91]. The lipid-binding site included the keto carbonyl group of spheroidenone and a Glu residue located two residues from the N-terminus of the -subunit. When changed to Ala by site-directed mutagenesis, a reduction of the phosphatidylethanolamine level in the lipid boundary phase was observed together with its replacement by phosphatidylcholine. ICM from the strain harboring the mutated protein retained a normal spherical vesicular morphology, but by exchanging the carotenoid from spheroidenone to neurosporene resulted in tubular membrane invagination in this strain. It was suggested that

24

ACCEPTED MANUSCRIPT this specific Glu residue as well as the nature of the carotenoid polar head group govern interactions with the surrounding phospholipid that together with LH2 participate in the ICM

PT

vesiculation process. This led to the conclusion that the preferential binding and packing of

RI

phospatidylethanolamine at the LH2-lipid interface is essential for maintaining ICM morphology.

SC

4. Dynamics of the developing ICM as examined by fluorescence emission analyses

NU

4.1 Bacteriochlorophyll a-based fluorescence induction/relaxation kinetics Fluorescence induction/relaxation measurements have served as a valuable method for

MA

assessment of the functional absorption cross-section, the charge separation and electron transport capabilities of anoxygenic phototropic bacteria [92-97]. While light-induced absorption

TE

D

changes for monitoring electron transfer reactions have largely prevailed for the evaluation of the physiological status of the purple anoxygenic phototrophs, several additional functional

AC CE P

parameters can be conveniently assessed from the analysis of the fluorescence emitted when RC phototraps are photooxidized and effectively become ―closed‖ during the fluorescence induction phase, and open again during the fluorescence relaxation phase. These include: (i) the quantum yield of the primary charge separation; (ii) the rates of RC electron transfer turnover; (iii) the functional absorption cross-section and connectivity of the photosynthetic apparatus (Fig. 7). Moreover, fluorescence spectroscopy is a highly sensitive technique for examining membrane dynamics in a component specific manner with both intrinsic and extrinsic fluorophores. The pioneering discovery of the relation between the fluorescence yield and the fraction of closed ―RC‖ phototraps by Duysens [96] has served as the basis for the current understanding of how excitation energy transfer, via inductive resonance, serves to couple the large LH antenna complement to multiple RC. When a RC is closed and, providing that little energy is lost via

25

ACCEPTED MANUSCRIPT non-radiative dissipative pathways, excitations can visit neighboring open RCs and ultimately undergo trapping. A structural basis for this concept has recently been borne out by AFM

PT

topographs of native Rba. sphaeroides membranes [13,14](Fig. 1A), in which many RC-LH1 core complexes are arranged in multiple rows of dimeric structures, facilitating the migration of

RI

an LH1 excitation along a series of dimers until an open RC is found. Analysis of the induction

SC

and relaxation characteristics of the fluorescence emitted during these trapping events is the basis

NU

for determining the various physiological parameters that interrogate the functional status of the photosynthetic apparatus.

MA

The variable fluorescence signal arising from the RC phototraps reflects the redox status of the RC, where a low fluorescence yield indicates an open RC (uncharged, P870 capable of

TE

+

D

performing photochemistry) and a high fluorescence yield indicates a fully closed RC (P870 photooxidized to P870 , charged RC transiently nonfunctional). The difference between the

AC CE P

minimal fluorescence (F0) and the maximal fluorescence (FM) yields upon closing of the RC, provides an estimate of the quantum yield of the primary charge separation, obtained from the FV/FM ratio, where variable fluorescence FV = FM-F0. In addition, analysis of the induction kinetics also provides an estimate of the functional absorption cross-section (σ) of the LH complexes, as well as their connectivity (p), the latter reflected by the sigmoidicity of the fluorescence induction curve. The fluorescence relaxation kinetics arise from the reopening of the RC as governed by the RC electron transfer turnover rate (τET). Importantly, the kinetics and energetics of the RC donor and acceptor side electron transfer reactions determine the nature of the fluorescence relaxation phase [43]. As noted by van Grondelle [97], three different RC redox states are known to cause substantially different fluorescence yields during the induction phase: (i) fully active ―open‖ RCs (oxidized P870, QA, QB) having a low fluorescence yield; (ii) closed

26

ACCEPTED MANUSCRIPT -

RCs which have twice the yield of open RCs if QA is reduced to the semiquinone anion (QA. ); (iii) closed RCs which have ~3-times the open RC yield when the primary donor is fully +

PT

oxidized to the P870 .

RI

Kato et al. [98] have recently noted that while Chl fluorescence techniques can serve as a

SC

highly accessible alternative method for monitoring the charge separation and O2 evolution activities of PSII in oxygenic phototrophs, fluorescence only indirectly assesses such parameters

NU

as RC charge, and is dependent on excitation light intensity and wavelength1 and whether the

MA

sample is dispersed in solution or in the solid state. Moreover, imprecise spectroscopic interpretations can arise from proteins that remain spectroscopically active after losing their

D

functional activities. Although these caveats may also apply to BChl fluorescence emission of

TE

purple anoxygenic phototrophs as assessed in induction/relaxation measurements, the results of our fluorescence kinetic analyses are largely corroborated by related structural and functional

AC CE P

correlates: (i) FV/FM ratios are in the range of 0.7-0.8, close to the ~0.65 obtained in Rbl. acidophilus in the presence of the QB-site inhibitor terbutryn [95]; (ii) fluorescence induction curves display essentially exponential kinetics consistent with single turnover behavior (Fig. 7); (iii) FM levels obtained by single turnover flashes were essentially matched by those arising from multiple turnover pulses on a tens of millisecond timescale; (iv) values obtained for functional absorption cross-sections ranged from 28-130 Å2 [23,29], corresponding closely to the range of BChl molecules/RC (30-120) which make up the photosynthetic unit. The lower value, 1Koblízek

et al. (29) observed a dependence of excitation wavelength on functional absorption cross-section values, which for steady-state Rba. sphaeroides cells excited at 470 and 795 nm were ~55 Å2 for 470 nm and ~65 Å2 for 795 nm. This can be accounted for by differences in energy transfer efficiency from carotenoids to LH1 BChl, which in the LH1 complex isolated by lithium dodecyl native electrophoresis [99] yielded a value of 61 % [100], accounting for the lower values for 470 nm, while a higher efficiency of B800  B850 energy transfer accounts for the higher 795 nm value. 27

ACCEPTED MANUSCRIPT representing the number of LH1 BChl molecules per RC, is close to that of 32 obtained in Rba. sphaeroides by spectral deconvolution coupled with RC photooxidation measurements [101],

PT

and from fluorescence induction in Rbl. acidophilus containing both the B800-850 and B800-820

RI

LH2 antenna complexes [95]. We believe that together, these data attest to the validity of near-IR fluorescence induction/relaxation analyses for monitoring selected photosynthetic activities of

SC

the purple anoxygenic phototrophs.

NU

It has been widely proposed that in the PSII RC of oxygenic phototrophs, variable fluorescence mainly reflects the redox status of primary stable electron acceptor QA (RC acceptor

MA

side)[102]. As noted above, recent reports [42,43] on the nature of the fluorescence relaxation phase in Rba. sphaeroides suggest a major role for the primary P870 electron donor (c-type

TE

D

cytochrome), depending upon the type of RC association of the donor cytochrome in the organism under examination. Accordingly, Asztalos et al. [43], using intact cells of Rba.

AC CE P

sphaeroides, Rhodospirillum rubrum and Rubriviax gelatinosus, showed that laser diode excitation at 808 nm yielded complex, multi-expontential fluorescence decay kinetics. Longer excitations were correlated with the slowing of the relaxation phase, which was determined by the redox status, size and accessibility of the reduced cytochrome c2 and quinone pools for bacterial strains limited on either the donor or acceptor side of the RC, respectively. In Rba. sphaeroides and Rsp. rubrum, in which mobile cytochromes c2 serves as the immediate P870 electron donor, relaxation is preferentially controlled by the re-reduction rate of the oxidized RC by the reduced cytochrome c2, and the acceptor side has only a minor role. For Rba. sphaeroides, the multiphasic P+ reduction kinetics have been explained by a proximal–distal cytochrome c2 binding model [103-105] where the fastest component (~1 s) reflects tight binding at the proximal site. The origin of the second phase (20-200s) is less clear and may result from

28

ACCEPTED MANUSCRIPT cytochrome c2 binding at the distal site or the cytochrome may move from other binding sites. The third phase (100s-10 ms) is thought to arise from the shuttling of cytochrome c2 between

PT

the RC and cytochrome bc1 complexes. The close correlation between the equilibrium redox +

RI

titration on the RC donor side and the fluorescence relaxation (43) further suggested that P

SC

reduction acts as the rate-limiting step the Rba. sphaeroides RC re-opening process. On the other hand, in Rvx. gelatinosus where the cytochrome subunit is physically attached to

NU

the RC, the acceptor side determines the fluorescence relaxation kinetics with the rate constant of -

MA

350 s—1 for QA interquinone electron transfer as the determining factor. Importantly, an analysis of the nature of the relaxation phase as a means of assessing membrane dynamics has

D

recently been applied to Rba. sphaeroides cells undergoing different developmental regimens

TE

(23,24,42), as will be described below.

4.2 Alterations in membrane dynamics as assessed by fluorescence induction/relaxation analyses

AC CE P

A near-IR fast-repetition rate fluorescence analysis2 by Koblízek et al. [23] in steady-state Rba. sphaeroides cells undergoing gratuitous induction of ICM formation at reduced oxygen tension in the dark was the first to suggest that ICM development was accompanied by major changes in membrane dynamics. A linear relation was observed between the rise in the 795

nm

and the elevations in the LH2/RC; this was expected since both parameters reflect the preferential insertion of LH2 into the membrane, while the LH1/RC remained constant, making up the core of the photosynthetic unit. Moreover, a correlation between the growth of the LH2 antenna and the gradual slowing of RC electron transfer turnover was observed. The

ET

was

relatively fast (0.8-1.6 ms) following aerobic growth and during the first 3 h at low aeration, but 2

Although for these initial studies, a near-IR fast repetition rate fluorometer [106] was employed in which the fluorescence signal was excited by a sequence of submicrosecond flashes rather

29

ACCEPTED MANUSCRIPT than a single 144 s single turnover flash as in the subsequent in the subsequent FIRe analyses, the results from both systems have proved to be essentially the same.

PT

a decrease to 7 ms was observed after 13 h, occurring in parallel with the rise in functional absorption cross-section. This was attributed to an imposition of constraints on electron transfer

RI

as the membrane bilayer becomes densely packed with accumulating LH2 rings. This possibility

SC

is supported further by fluorescence induction/relaxation measurements in cells undergoing adaptation from high to low light intensity. The rates of RC electron transfer turnover and

NU

reoxidation of UQ pool were decreased by ~4- and 3-fold, respectively, over 4 days [23].

MA

Moreover, the slowing of ET was directly related to the increase in the functional absorption cross-section (450 nm ranged from ~70 Å2 at outset to 130 Å2 after 4 days) [23,24]. Again, an

D

essentially linear relation was observed between the size of the absorption cross-section and the

TE

LH2:LH1 molar ratios. Thus, when applied to the monitoring of changes in photosynthetic

AC CE P

activities during ICM development, taken together with AFM surface views of the alterations in membrane structure (13,14), fluorescence induction/relaxation measurements have served as a valuable probe of the membrane dynamics accompanying the membrane remodeling process (23,27).

While we initially attributed the slowing of the ET as a function of increasing LH2 levels to the imposition of constraints upon free diffusion of ubiquinone redox species between the RC and cytochrome bc1 complex (RC acceptor side kinetics)[23,24] as discussed above, a somewhat more complicated picture has emerged from the studies on Rba. sphaeroides by Kis et al. [42]. In cells adapting from aerobic to phototrophic growth conditions under constant illumination, donor electron transfer reactions dominated the relaxation kinetics after a single-turnover flash, while contributions on the acceptor side were enhanced only after multiple turnover flashes. It was

30

ACCEPTED MANUSCRIPT concluded that the slowing of electron transfer turnover results from changes on the RC donor side in which an increased distance for electron flow occurs in the cytochrome c2 periplasmic

PT

diffusion pathway between the RC and cytochrome bc1 complexes. This accounts for the initial

RI

relaxation kinetics when the developed (mature) membrane becomes more densely packed with LH2 rings and/or the distal RC position [103-105] becomes less accessible to cytochrome c2.

SC

Moreover, Kis et al. [42] showed that the slowed rate of electron transfer during ICM

NU

development did not result from changes in cytochrome c2 pool size. On the other hand, slow reoxidation of the UQ pool reflected contributions of the acceptor side, which is enhanced after

MA

long excitation (multiple turnovers). This is also compatible with our results [23,24], which were interpreted to show slowing of UQ pool oxidation with increasing LH2 levels after multiple

TE

D

turnover flashes. However, AFM images of Rba. sphaeroides LH2-only membranes have formed the basis for a functional model of highly curved native membranes, which holds that these

AC CE P

structures would permit the free diffusion of electron carriers through the LH2 domains, while still maintaining efficient excitation energy transfer [107]. Clearly, more work needs to be done on the question of whether quinone redox species flow is impeded by the presence of hexagonally packed LH2-only domains. 4.3. Searching for a mechanism explaining the decreases in reaction center electron transfer turnover rate during greening and adaptation to low light intensity A second question arises concerning the growth of the Rba. sphaeroides ICM during adaptation to reduced oxygen tension or low light intensity and the distance traveled by cytochrome c2 as predicted from the fluorescence induction studies of Kis et al. [42]. The intracelleular status of the Rba. sphaeroides ICM is now a matter of intense debate [16,108,109]. The cryo-electron tomography images of Tucker et al. [16] revealed a complex scenario for the

31

ACCEPTED MANUSCRIPT ultrastructure assumed by the ICM. While some ICM vesicles were observed without connections to other structures, others were located nearer to the CM, frequently forming

PT

interconnections with physical association to the CM, providing apparent access to the

RI

periplasm. Some nearly spherical single invaginations, with ―necks‖ attaching them to the CM were also seen, along with small CM indentations, representing the ICM precursor membrane

SC

isolated in the upper pigmented band in sucrose density gradients of cell-free extracts. Thus,

NU

fully detached ICM vesicles, possessing all the machinery for converting light energy into ATP, regarded as bacterial membrane organelles, only represent a part of the picture, for which the

MA

tomography methodology does not permit quantification of their relative content. A functional approach to the question of whether an ICM continuum exists with free access to

TE

D

the periplasmic space has been taken recently [108], where the ionophore gramicidin was employed to investigate the relative size of the electroosmotic units in intact cells, isolated

AC CE P

chromatophores and spheroplasts, lacking an outer membrane. In intact cells and in spheroplasts, the half-time of the decay induced by a single channel was ~6 ms and ~23 ms, respectively, as measured from the carotenoid electrochromic shift, which was three to four orders of magnitude slower that in isolated chromatophores. This was interpreted to mean that that the area of functionally active membrane in cells or spheroplasts is at least three orders of magnitude greater than that in individual chromatophore vesicles, which at most could contribute 10% of the active photosynthetic membrane in the cell. This is to be compared with the results of Prince et al. [110] who showed that ≤85 % of the cytochrome c2 was released upon spheroplast formation and thus an upper limit for free chromatophores would be ~15%; however, this is likely to amount to more at low light intensities. It was emphasized [108] that a single internal membrane system with connections to the periplasmic space may provide significant advantages in renewing the

32

ACCEPTED MANUSCRIPT photosynthetic apparatus and in the reallocation of the components shared with additional bioenergetic pathways.

PT

A more recent report on the cryo-electron tomography of the ICM from Rba. sphaeroides has

RI

appeared [109] in which the results were interpreted to show that ICM vesicles form a continuous reticulum rather than existing as discrete vesicular structures. The possibility must be

SC

considered that the different 3-D ICM arrangements are due to strain difference, since the Ga

NU

strain used in this study was derived from the 2.4.1 strain in which the yield of ICM vesicles is much lower than in strain NCIB8253 [27]. Most of the photopigment is retained in an upper-

MA

pigmented band in sucrose gradients, arising from CM invagination sites. This means that the internal membrane invagination and budding process may be limited in the 2.4.1 and Ga strains.

TE

D

In this study [109], the cytochrome bc1 complex was localized in the ICM, through immunoelectron microscopy using immuno-gold labeling, with antibodies raised against peptides

AC CE P

representing putative, well conserved cytoplasmic loops of cytochrome b from Rsp. rubrum. Labeling was mostly found on small membrane fragments and in the ―neck‖ regions protruding from the ICM vesicles suggesting that that the complexes localize to the fragile membrane regions that interconnect the chromatophores in the network. These fragile regions were thought to arise from tubular structures reconstructed from tomographic data, seen lying between ICM vesicles. Further, it was proposed that the bc1 complexes within these regions are surrounded by RC-LH1 complexes, bordered by arrays of the peripheral LH2 antenna complex. The localization of the cytochrome bc1 complex in the native ICM has also been examined recently by Cartron et al. [111]. EMs of negatively-stained Rba. sphaeroides chromatophores in which the C-terminus of cytochrome c1 contained a His10-tag that was labeled with gold nanobeads, showed that the majority of the labeled complexes were dimeric (Fig. 8A,B). The bc1

33

ACCEPTED MANUSCRIPT complexes appeared to be adjacent to RC-LH1 complexes, as seen in AFM topographs of goldlabeled chromatophores (Fig. 8C). While this provides evidence that restricted electron transfer

PT

domains might exist, these were not seen as fixed RC-LH1-PufX-cytochrome bc1

RI

supercomplexes. Instead, cytochrome bc1 complexes sit adjacent to the core complexes in disordered areas, not necessarily in direct contact with the majority of the RC-LH1-PufX

SC

complexes (Fig. 8). These findings were combined with a quantitative MS analysis in which the

NU

levels of the RC, cytochrome bc1 complex, ATP synthase, and cytochrome aa3 and cbb3 complexes were determined as a basis for constructing an atomic resolution model of a 50-nm

MA

chromatophore vesicle. The modeled vesicle contained 67 LH2 complexes, 11 LH1–RC dimers, 2 RC–LH1 monomers, 4 cytochrome bc1 dimers and 2 ATP synthases. Through simulation of the

TE

D

energy, electron and proton transfer processes, the calculated half-maximum ATP turnover rate suggested that this photosynthetic unit architecture is optimized for accommodating to low light

AC CE P

intensities.

Since the current atomic-level structural model for a 50-nm low-light chromatophore has four dimeric cytochrome bc1 complexes in close arrangement with the RC-LH1 complex, with seven RC-LH1 at some distance, the average distances over the internal vesicle surface for diffusion of cytochrome c2 between cytochrome bc1 and the RC is 10–30 nm. This may account for the observed slowing of the RC donor pathway, but localization of the bc1 complex in high-light chromatophores in which the LH2 content is reduced, will be essential for confirming this possibility. Acknowledgements We thank Prof. Peter Lobel and Dr. Haiyan Zheng for conducting the proteomics analyses and Dr. Kamil Woronowicz, Dr. Raoul Frese, Oluwatobi B. Olubanjo and Daniel Sha for

34

ACCEPTED MANUSCRIPT participating in the studies performed in my laboratory. Work in the author’s laboratory was supported by grants from the U. S. Department of Energy, Grant No. DE-FG02-08ER15957

PT

from the Chemical Sciences, Geosciences and Biosciences Division, Office of Basic Energy

AC CE P

TE

D

MA

NU

SC

RI

Sciences and the National Science Foundation (Subaward No. 12-764).

35

ACCEPTED MANUSCRIPT References [1] C.N. Hunter, F. Daldal, M.C. Thurnauer, J.T. Beatty (Eds.), The Purple Phototrophic

PT

Bacteria, Springer, Netherlands, Dordrecht, 2008.

RI

[2] J. Deisenhofer, H. Michel, The photosynthetic reaction center from the purple bacterium Rhodopseudomonas viridis, Science 245 (1989) 1463–1473.

SC

[3] J.P. Allen, G. Feher, T.O. Yeates, H. Komiya, D.C. Rees, Structure of the reaction center

NU

from Rhodobacter sphaeroides R-26: the cofactors, Proc. Natl. Acad. Sci. U S A. 84 (1987) 5730-5734.

MA

[4] J.P. Allen, G. Feher, T.O. Yeates, H. Komiya, D.C. Rees, Structure of the reaction center from Rhodobacter sphaeroides R26: The protein subunits, Proc. Natl. Acad. Sci. U.S.A. 84

TE

D

(1987) 6162–6166.

[5] C.H. Chang, D. Tiede, J. Tang, U. Smith, J. Norris, M. Schiffer, Structure of

AC CE P

Rhodopseudomonas sphaeroides R-26 reaction center, FEBS Lett 205 (1986) 82–86. [6] G. McDermott, S.M. Prince, A.A. Freer, A.M. Hawthornthwaite-Lawless, M.Z. Papiz, R.J. Cogdell, N.W. Isaacs, Crystal structure of an integral membrane light-harvesting complex from photosynthetic bacteria, Nature 374 (1995) 517–521. [7] J. Koepke, X.C. Hu, C. Muenke, K. Schulten, H. Michel, The crystal structure of the lightharvesting complex II (B800-850) from Rhodospirillum molischianum, Structure 4 (1996) 581–597. [8] A.W. Roszak, T.D. Howard, J. Southall, A.T. Gardiner, C.J. Law, N.W. Isaacs, R.J. Cogdell, Crystal structure of the RC-LH1 core complex from Rhodopseudomonas palustris, Science 302 (2003) 1969–1972. [9] S. Scheuring, D. Levy, J.L. Rigaud, Watching the components of photosynthetic bacterial

36

ACCEPTED MANUSCRIPT membranes and their in situ organisation by atomic force microscopy, Biochim. Biophys. Acta 1712 (2005) 109-127.

PT

[10] J.N. Sturgis, J.D. Tucker, J.D. Olsen, C.N. Hunter, R.A. Niederman, Atomic force

RI

microscopy studies of native photosynthetic membranes, Biochemistry 48 (2009) 36793698.

SC

[11] S. Scheuring, J.N. Sturgis, Atomic force microscopy of the bacterial photosynthetic

NU

apparatus: plain pictures of an elaborate machinery, Photosynth. Res. 102 (2009) 197-211. [12] L.-N. Liu, S. Scheuring, Investigation of photosynthetic membrane structure using atomic

MA

force microscopy, Trends Plant Sci. 18 (2013) 277-286. [13] S. Bahatyrova, R.N. Frese, C.A. Siebert, K.O. van der Werf, R. van Grondelle, R.A.

TE

D

Niederman, P.A. Bullough, C. Otto, J.D. Olsen, C.N. Hunter, The native architecture of a photosynthetic membrane, Nature 430 (2004) 1058–1062.

AC CE P

[14] P.G. Adams, C.N. Hunter, Adaptation of intracytoplasmic membranes to altered light intensity in Rhodobacter sphaeroides, Biochim Biophys Acta 1817 (2012) 1616–1627. [15] S. Scheuring, J.N. Sturgis, Chromatic adaptation of photosynthetic membranes. Science 309 (2005) 484–487.

[16] J.D. Tucker, C.A. Siebert, M. Escalante, P.G. Adams, J.D. Olsen, C. Otto, D.L. Stokes, C.N. Hunter, Membrane invagination in Rhodobacter sphaeroides is initiated at curved regions of the cytoplasmic membrane, then forms both budded and fully detached spherical vesicles, Mol. Microbiol. 76 (2010) 833–847. [17] J.N. Sturgis, C.N. Hunter, R.A. Niederman, Spectra and extinction coefficients of near infrared absorption bands in membranes of Rhodobacter sphaeroides mutants lacking lightharvesting and reaction center complexes, Photochem. Photobiol. 48 (1988) 243–247.

37

ACCEPTED MANUSCRIPT [18] S. Scheuring, J.N. Sturgis, Dynamics and diffusion in photosynthetic membranes from Rhodospirillum photometricum, Biophys. J. 91 (2006) 3707–3717.

PT

[19] M.K. Sener, J.D. Olsen, C.N. Hunter, K. Schulten, Atomic-level structural and functional

RI

model of a bacterial photosynthetic membrane vesicle. Proc. Natl. Acad. Sci. U.S.A. 104 (2007) 15723–15728.

SC

[20] X. Zeng, J.H. Roh, S.J. Callister, C.L. Tavano, T.J. Donohue, M.S. Lipton, S. Kaplan,

NU

Proteomic characterization of the Rhodobacter sphaeroides 2.4.1 photosynthetic membrane: identification of new proteins, J. Bacteriol. 189 (2007) 7464–7474.

MA

[21] K. Woronowicz, R.A. Niederman, Proteomic analysis of the developing intracytoplasmic membrane in Rhodobacter sphaeroides during adaptation to low light intensity, Adv. Exp.

TE

D

Med. Biol. 675 (2010) 161–178.

[22] G.M. D'Amici, S. Rinalducci, L. Murgiano, F. Italiano, L. Zolla, Oligomeric

AC CE P

characterization of the photosynthetic apparatus of Rhodobacter sphaeroides R26.1 by nondenaturing electrophoresis methods, J. Proteomic Res. 9 (2010) 192–203. [23] K. Woronowicz, D. Sha, R.N. Frese, R.A. Niederman, The accumulation of the light harvesting 2 complex during remodeling of the Rhodobacter sphaeroides intracytoplasmic membrane results in a slowing of the electron transfer turnover rate of photochemical reaction centers, Biochemistry 50 (2011) 4819–4829. [24] K. Woronowicz, D. Sha, R.N. Frese, J.N. Sturgis, V. Nanda, R.A. Niederman, The effects of protein crowding in bacterial photosynthetic membranes on the flow of quinone redox species between the photochemical reaction center and the ubiquinol cytochrome c2 oxidoreductase, Metallomics 3 (2011) 765–774. [25] K. Woronowicz, O.B. Olubanjo, H.C. Sung, J.L. Lamptey, R.A. Niederman, Differential

38

ACCEPTED MANUSCRIPT assembly of polypeptides of the light-harvesting 2 complex encoded by distinct operons during acclimation of Rhodobacter sphaeroides to low light intensity, Photosynth. Res. 108

PT

(2011) 201–214.

RI

[26] P.J. Jackson, H.J. Lewis, J.D. Tucker, C.N. Hunter, M.J. Dickman, Quantitative proteomic analysis of intracytoplasmic membrane development in Rhodobacter sphaeroides, Mol.

SC

Microbiol. 84 (2012) 1062–1067.

NU

[27] K. Woronowicz, J.W. Harrold, J.M. Kay, R.A. Niederman, Structural and functional proteomics of intracytoplasmic membrane assembly in Rhodobacter sphaeroides, J. Mol.

MA

Microbiol. Biotechnol. 23 (2013) 48–62.

[28] R.A. Niederman, Membrane development in purple photosynthetic bacteria in response to

TE

D

alterations in light intensity and oxygen tension, Photosynth. Res. 116 (2013) 333–348. [29] M. Koblízek, J.D. Shih, S.I. Breitbart, E.C. Ratcliffe, Z.S. Kolber, C.N. Hunter, R.A.

AC CE P

Niederman, Sequential assembly of photosynthetic units in Rhodobacter sphaeroides as revealed by fast repetition rate analysis of variable bacteriochlorophyll a fluorescence, Biochim. Biophys. Acta 1706 (2005) 220–231. [30] G. Cohen-Bazire, W.R. Sistrom, R.Y. Stanier, Kinetic studies of pigment synthesis by nonsulfur purple bacteria, J. Cell Comp. Physiol. 49 (1956) 25–68. [31] J. Takemoto, J. Lascelles, Coupling between bacteriochlorophyll and membrane protein synthesis in Rhodopseudomonas sphaeroides, Proc. Natl. Acad .Sci. USA 70 (1973) 799– 803. [32] K. Woronowicz, O.B. Olubanjo, D. Sha, J.M. Kay, R.A. Niederman, Effects of the protonophore carbonyl-cyanide m-chlorophenylhydrazone on intracytoplasmic membrane assembly in Rhodobacter sphaeroides, Biochim. Biophys. Acta 1847 (2015) 1119-1128.

39

ACCEPTED MANUSCRIPT [33] I. Wittig, M. Karas, H. Schägger, High resolution clear native electrophoresis for in-gel functional assays and fluorescence studies of membrane protein complexes, Mol. Cell.

PT

Proteomics 6 (2007) 1215-1225.

RI

[34] C. Mackenzie, M. Choudhary, F.W. Larimer, P.F. Predki, S. Stilwagen, J.P. Armitage, R.D. Barber, T.J. Donohue, J.P. Hosler, J.E. Newman, J.P. Shapleigh, R.E. Sockett, J. Zeilstra-

SC

Ryalls, S. Kaplan, The home stretch, a first analysis of the nearly completed genome of

NU

Rhodobacter sphaeroides 2.4.1, Photosynth. Res. 70 (2001) 19-41. [35] R.L. Tatusov, E.V. Koonin, D.J. Lipman, A genomic perspective on protein families,

MA

Science 278 (1997) 631-637.

[36] P.G. Heytler, Uncoupling of oxidative phosphorylation by carbonyl cyanide

TE

D

phenylhydrazones. I. Some characteristics of m-Cl-CCP action on mitochondria and chloroplasts, Biochemistry 2 (1963) 357-361.

AC CE P

[37] J. Cunarro, M.W. Weiner, Mechanism of action of agents which uncouple oxidative phosphorylation: direct correlation between proton-carrying and respiratory-releasing properties using rat liver mitochondria, Biochim. Biophys. Acta 387 (1975) 234-240. [38] R. Dierstein, G. Drews, Effect of uncoupler on assembly pathway for pigment binding protein of bacterial photosynthetic membranes, J. Bacteriol. 168 (1986) 167–172. [39] J.N. Sturgis, R.A. Niederman, Organization and assembly of light-harvesting complexes in the purple bacterial membrane, in: C.N. Hunter, F. Daldal, M.C. Thurnauer, J.T. Beatty (Eds.), The Purple Phototrophic Bacteria, Springer, Netherlands, Dordrecht 2008, pp. 253– 273. [40] G. Drews, Formation of the light-harvesting complex I (B870) of anoxygenic phototrophic purple bacteria, Arch. Microbiol. 166 (1996) 151–159.

40

ACCEPTED MANUSCRIPT [41] M. Meryandini, G. Drews, Import and assembly of the α- and β-polypeptides of the lightharvesting complex I (B870) in the membrane system of Rhodobacter capsulatus

PT

investigated in an in vitro translation system, Photosynth. Res. 47 (1996) 21–31.

RI

[42] M. Kis, E. Asztalos, G. Sipka, P. Maróti, Assembly of photosynthetic apparatus in Rhodobacter sphaeroides as revealed by functional assessments at different growth phases

SC

and in synchronized and greening cells, Photosynth. Res. 122 (2014) 261-273.

NU

[43] E. Asztalos, G. Sipka, P. Maróti, Fluorescence relaxation in intact cells of photosynthetic bacteria: donor and acceptor side limitations of reopening of the reaction center. Photosynth.

MA

Res. 124 (2015) 31-44.

[44] N. Nouwen, A.J. Driessen, SecDFYajC forms a heterotetrameric complex with YidC, Mol.

TE

D

Microbiol. 44 (2002) 1397-1405.

[45] N. Tavernarakis, M. Driscoll, N.C. Kyrpides, The SPFH domain: implicated in regulating

AC CE P

targeted protein turnover in stomatins and other membrane-associated proteins, Trends Biochem. Science 24 (1999) 425-427. [46] K. Ito, Y. Akiyama, Cellular functions, mechanism of action, and regulation of FtsH protease. Annu. Rev. Microbiol. 59 (2005) 211–231. [47] E. van Bloois, H.L. Dekker, L. Froderberg, E.N.G. Houben, M.L. Urbanus, C.G. de Koster, J.-W. de Gier, J. Luirink, Detection of cross-links between FtsH, YidC, HflK/C suggests a linked role for these proteins in quality control upon insertion of bacterial inner membrane proteins, FEBS Lett. 582 (2008) 1419-1424. [48] J. Luirink, I. Sinning, SRP-mediated protein targeting: structure and function revisited, Biochim. Biophys. Acta 1694 (2004) 17-35.

41

ACCEPTED MANUSCRIPT [49] M. Paetzel, R.E. Dalbey, N.C. Strynadka, Crystal structure of a bacterial signal peptidase in complex with a beta-lactam inhibitor, Nature 396 (1998) 186-190.

PT

[50] S.E. Lang, F.E. Jenney, F. Daldal, Rhodobacter capsulatus CycH: a bipartite gene product

RI

with pleiotropic effects on the biogenesis of structurally different c-type cytochromes, J. Bacteriol. 178 (1996) 5279-5290.

SC

[51] J. Bachmann, B. Bauer, K. Zwicker, B. Ludwig, O. Anderka, The Rieske protein from

NU

Paracoccus denitrificans is inserted into the cytoplasmic membrane by the twin-arginine translocase, FEBS J. 273 (2006) 4817–4830.

MA

[52] C.S. Young, J.T. Beatty, Multi-level regulation of purple bacterial light-harvesting complexes, in: B.R. Green W.W. Parson, (Eds.) Light-Harvesting Antennas In

TE

D

Photosynthesis, Kluwer Academic Publishers, Dordrecht, The Netherlands, (2003), pp. 449470.

AC CE P

[53] J.D. Olsen, P.G. Adams, P.J. Jackson, M.J. Dickman, P. Qian, C.N. Hunter, Aberrant assembly complexes of the reaction center light-harvesting 1 PufX (RC-LH1-PufX) core complex of Rhodobacter sphaeroides imaged by atomic force microscopy, J. Biol. Chem. 289 (2014) 29927-2996.

[54] J. Komenda, V. Reisinger, B.C. Müller, M. Dobáková B. Granvogl, L.A. Eichacker, Accumulation of the D2 protein is a key regulatory step for assembly of the photosystem II reaction center complex in Synechocystis PCC 6803, J. Biol. Chem. 279, 48620-48629. [55] M. Dobáková, R. Sobotka, M. Tichý, J. Komenda, Psb28 protein is involved in the biogenesis of the photosystem II inner antenna CP47 (PsbB) in the cyanobacterium Synechocystis sp. PCC 68031, Plant Physiol. 149 (2009) 1076-1086. [56] J. Komenda, J. Knoppová, J. Kopečná, R. Sobotka, P. Halada, J.F. Yu, J. Nickelsen, M.

42

ACCEPTED MANUSCRIPT Boehm, P.J. Nixon, The Psb27 assembly factor binds to the CP43 complex of photosystem II in the cyanobacterium Synechocystis sp. PCC 68031, Plant Physiol. 158 (2012) 476–486.

PT

[57] J.W. Chidgey, M. Linhartová. J. omenda, P.J. Jackson, M. J. ickman, . P. Canniffe, P.

RI

on k, J. Piln , C.N. Hunter, R. Sobotka, A cyanobacterial chlorophyll synthase-HliD complex associates with the Ycf39 protein and the Yidc/Alb3 insertase, Plant Cell 26 (2014)

SC

1267–1279.

NU

[58] A. Einhauer, A. Jungbauer, The FLAG peptide, a versatile fusion tag for the purification of recombinant proteins, J. Biochem. Biophys. Methods 49 (2001) 455–465.

MA

[59] J. Kim, P.G. Klein, J.E. Mullet, Synthesis and turnover of photosystem II reaction center protein D1. Ribosome pausing increases during chloroplast development, J. Biol. Chem. 269

TE

D

(1994) 17918–17923.

[60] J.C. Samuelson, M. Chen, F. Jiang, I. Moller, M. Wiedmann, A. Kuhn, G.J. Phillips, R.E.

AC CE P

Dalbey, YidC mediates membrane protein insertion in bacteria, Nature 406 (2000) 637-641. [61] I. Sachelaru, N.A. Petriman, R. Kudva, P. Kuhn, T. Welte, B. Knapp, F. Drepper, B. Warscheid, H.-G. Koch, YidC occupies the lateral gate of the SecYEG translocon and is sequentially displaced by a nascent membrane protein, J. Biol. Chem. 288 (2013) 1629516307.

[62] L. Zhu, H.R. Kaback, R.E. Dalbey, YidC protein, a molecular chaperone for LacY protein folding via the SecYEG protein machinery, J. Biol. Chem. 288 (2013) 28180-28194. [63] L. Yi, F. Jiang, M. Chen, B. Cain, A. Bolhuis, R.E. Dalbey, YidC is strictly required for membrane insertion of subunits a and c of the F1F0ATP synthase and SecE of the SecYEG translocase, Biochemistry 42 (2003) 10537-10544. [64] M. Van Der Laan, M.L. Urbanus, C.M. Ten Hagen-Jongman, N. Nouwen, B. Oudega, N.

43

ACCEPTED MANUSCRIPT Harms, A. J. Driessen, J. Luirink, A conserved function of YidC in the biogenesis of respiratory chain complexes, Proc. Natl. Acad. Sci. U.S.A. 100 (2003) 5801-5806.

PT

[65] N. Nouwen, A.J. Driessen, SecDFYajC forms a heterotetrameric complex with YidC, Mol.

RI

Microbiol. 44 (2002) 1397-1405.

[66] R.J. Schulze, J. Komar, M. Botte, W.J. Allen, S. Whitehouse, V.A. Gold, A. Lycklama, J.A.

SC

Nijehol, K. Huard, I. Berger, C. Schaffitzel, I. Collinson, Membrane protein insertion and

NU

proton-motive-force-dependent secretion through the bacterial holo-translocon SecYEGSecDF-YajC-YidC, Proc. Natl. Acad. Sci. USA 111 (2014) 4844-4849

MA

[67] C.N. Hunter, J.D. Tucker, R.A. Niederman, The assembly and organisation of photosynthetic membranes in Rhodobacter sphaeroides, Photochem. Photobiol. Sci. 4

TE

D

(2005) 1023-1027.

[68] R.A. Niederman, Eukaryotic behaviour of a prokaryotic energy transducing membrane:

AC CE P

fully detached vesicular organelles arise by budding from the Rhodobacter sphaeroides intracytoplasmic photosynthetic membrane, Mol. Microbiol. 76 (2010) 803-805. [69] C.E. Bauer, B.L. Marrs, Rhodobacter capsulatus puf operon encodes a regulatory protein (PufQ) for bacteriochlorophyll biosynthesis, Proc. Natl. Acad. Sci. U S A 85 (1988) 70747078.

[70] S. Fidai, S.B. Hinchigeri, T.J. Borgford, W.R. Richards, Identification of the PufQ protein in membranes of Rhodobacter capsulatus, J. Bacteriol. 176 (1994) 7244-7251. [71] S. Fidai, S.B. Hinchigeri, W.R. Richards, Association of protochlorophyllide with the PufQ protein of Rhodobacter capsulatus, Biochem. Biophys. Res. Commun. 200 (1994) 16791684.

44

ACCEPTED MANUSCRIPT [72] M. Aklujkar, R.C. Prince, J.T. Beatty, The PuhB protein of Rhodobacter capsulatus functions in photosynthetic reaction center assembly with a secondary effect on light-

PT

harvesting complex 1, J. Bacteriol. 187 (2005) 1334-1343.

RI

[73] M. Aklujkar, R.C. Prince, J.T. Beatty, The photosynthetic deficiency due to puhC gene deletion in Rhodobacter capsulatus suggests a PuhC protein-dependent process of

SC

RC/LH1/PufX complex reorganization, Arch. Biochem. Biophys. 454 (2006) 59-71.

NU

[74] M. Aklujkar, R.C. Prince, J.T. Beatty. The puhE gene of Rhodobacter capsulatus is needed for optimal transition from aerobic to photosynthetic growth and encodes a putative negative

MA

modulator of bacteriochlorophyll production, Arch. Biochem. Biophys. 437 (2005) 186-198. [75] J.H. Roh, W.E. Smith, S. Kaplan, Effects of oxygen and light intensity on transcriptome

TE

D

expression in Rhodobacter sphaeroides 2.4.1. Redox active gene expression profile, J. Biol. Chem. 279 (2004) 9146-9155.

AC CE P

[76] F. Francia, J. Wang, G. Venturoli, B.A. Melandri, W.P. Barz, D. Oesterhelt, The reaction center-LH1 antenna complex of Rhodobacter sphaeroides contains one PufX molecule which is involved in dimerization of this complex, Biochemistry 38 (1999) 6834–6845. [77] P. Qian, M.Z. Papiz, P.J. Jackson, A.A. Brindley, I.W. Ng, J.D. Olsen, M.J. Dickman, P.A. Bullough, C.N. Hunter, Three-dimensional structure of the Rhodobacter sphaeroides RCLH1-PufX complex: dimerization and quinone channels promoted by PufX, Biochemistry 52 (2013) 7575-7585. [78] W.P. Barz, A. Vermeglio, F. Francia, G. Venturoli, B.A. Melandri, D. Oesterhelt, Role of the PufX protein in photosynthetic growth of Rhodobacter sphaeroides. 2. PufX is required for efficient ubiquinone/ubiquinol exchange between the reaction center QB site and the cytochrome bc1 complex, Biochemistry 34 (1995) 15248-15258.

45

ACCEPTED MANUSCRIPT [79] R.N. Frese, J.D. Olsen, R. Branvall, W.H.J. Westerhuis, C.N. Hunter, R. van Grondelle, The long-range supra-organization of the bacterial photosynthetic unit: A key role for PufX,

PT

Proc. Natl. Acad. Sci. USA 97 (2000) 5197-5202.

RI

[80] R.N. Frese, C.A. Siebert, R.A. Niederman, C.N. Hunter, C. Otto, R. van Grondelle, The long-range organization of a native photosynthetic membrane, Proc. Natl. Acad. Sci. USA

SC

101 (2004) 17994-17999.

NU

[81] J. Abramson, I. Smirnova, V. Kasho G. Verner, H.R. Kaback, S. Iwata, Structure and mechanism of the lactose permease of Escherichia coli, Science 301 (2003) 610-615.

MA

[82] W.H.J. Westerhuis, J.N. Sturgis, E.C. Ratcliffe, C.N. Hunter, R.A. Niederman, Isolation, size estimates and spectral heterogeneity of an oligomeric series of light-harvesting 1

TE

D

complexes from Rhodobacter sphaeroides, Biochemistry 41 (2002) 8698-8707. [83] K. Sznee, L.I. Crouch, M.R. Jones, J.P. Dekker, R.N. Frese, Variation in supramolecular

AC CE P

organisation of the photosynthetic membrane of Rhodobacter sphaeroides induced by alteration of PufX, Photosynth. Res. 119 (2014) 243–256. [84] P. Qian, P.A. Bullough, C.N. Hunter, Three-dimensional reconstruction of a membranebending complex: the RC–LH1–PufX core dimer of Rhodobacter sphaeroides, J. Biol. Chem. 283 (2008) 14002–14011. [85] R.J. Pugh, P. McGlynn, M.R. Jones, C.N. Hunter, The LH1-RC core complex of Rhodobacter sphaeroides: interaction between components, time-dependent assembly, and topology of the PufX protein, Biochim. Biophys. Acta 1366 (1998) 301–316. [86]. L.-N. Liu, K. Duquesne, F. Oesterhelt, J.N. Sturgis, S. Scheuring, Forces guiding assembly of light-harvesting complex 2 in native membranes, Proc. Natl. Acad. Sci. USA 108 (2011) 9455–9459.

46

ACCEPTED MANUSCRIPT [87] C.N. Hunter, J.D. Pennoyer, J.N. Sturgis, D. Farrelly, R.A. Niederman, Oligomerization states and associations of light-harvesting pigment-protein complexes of Rhodobacter

PT

sphaeroides as analyzed by lithium dodecyl sulfate-polyacrylamide gel electrophoresis.

RI

Biochemistry 27 (1988) 3459–3467.

[88] P.J. Kiley, A. Varga, S. Kaplan, Physiological and structural analysis of light-harvesting

SC

mutants of Rhodobacter sphaeroides, J. Bacteriol. 170 (1988) 1103–1115.

NU

[89] J. Hsin, J. Gumbart, L. G. Trabuco, E. Villa, Pu Qian, C. Neil Hunter, K. Schulten, Protein-

Biophys. J. 97 (2009) 321–329.

MA

induced membrane curvature investigated through molecular dynamics flexible fitting

[90] D. Chandler, J. Hsin, C.B. Harrison, J. Gumbart, K. Schulten, Intrinsic curvature properties

D

of photosynthetic proteins in chromatophores. Biophys. J. 95 (2008.) 2822–2836.

TE

91 L. . wa, . Wegmann, . ru gger, F.T. Wieland, G. Wanner, P. Braun, Mutation of a

AC CE P

single residue, -glutamate-20, alters protein–lipid interactions of light harvesting complex II, Mol. Microbiol. 67((2008) 63–77. _ [92] K.L. Zankel, D.W. Reed, R.K. Clayton, Fluorescence and photochemical quenching in photosynthetic reaction centers, Proc. Natl. Acad. Sci. U. S. A. 61 (1968) 1243–1249. [93] C.N. Hunter, R. van Grondelle, N.G. Holmes, O.T.G. Jones, R.A. Niederman, Fluorescence yield properties of a fraction enriched in newly synthesized bacteriochlorophyll a–protein complexes from Rhodopseudomonas sphaeroides, Photochem. Photobiol. 30 (1979) 313– 316. [94] H.-W. Trissl, Antenna organization in purple bacteria investigated by means of fluorescence induction curves, Photosynth. Res. 47 (1996) 175– 185. [95] C.J. Law, R.J. Cogdell, H.-W. Trissl, Antenna organisation in the purple bacterium

47

ACCEPTED MANUSCRIPT Rhodopseudomonas acidophila studied by fluorescence induction, Photosynth. Res. 52 (1997) 157– 165.

PT

[96] L.N.M. Duysens, Transfer of light energy within the pigment systems present in

RI

photosynthesizing cells, Nature 168 (1951) 548.

[97] R. van Grondelle, Excitation energy transfer, trapping and annihilation in photosynthetic

SC

systems, Biochim. Biophys. Acta 811 (1985) 147-195.

NU

[98] M. Kato, J. Z. Zhang, N. Paul, E. Reisner, Protein film photoelectrochemistry of the water oxidation enzyme photosystem II, Chem. Soc. Rev. 43 (2014) 6485-6497.

MA

[99] R.M. Broglie, C.N. Hunter, P. Delepelaire, R.A. Niederman. N.-H. Chua, R.K. Clayton, Isolation and characterization of pigment-protein complexes of Rhodopseudomonas

TE

D

sphaeroides by lithium dodecyl sulfate/polyacrylamide gel electrophoresis, Proc. Natl. Acad. Sci. USA 77 (1980) 87-91.

AC CE P

[100] C.N. Hunter, R.A. Niederman, R.K. Clayton, Excitation energy transfer by antenna complexes isolated from Rhodopseudomonas sphaeroides by lithium dodecyl sulfate/polyacrylamide gel electrophoresis, in: G. Akoyunoglou, (Ed.) Photosynthesis, Structure and Molecular Organization of the Photosynthetic Apparatus, Vol. 3, Balaban, Philadelphia, (1981) pp. 539-545. [101] W.H.J. Westerhuis, J.W. Farchaus, R.A. Niederman, Altered spectral properties of the B875 light-harvesting pigment-protein complex in a Rhodobacter sphaeroides mutant lacking pufX, Photochem. Photobiol. 58 (1993) 460-463. [102] Z.S. Kolber, O. Prasil, P.G. Falkowski, Measurements of variable chlorophyll fluorescence using fast repetition rate techniques: defining methodology and experimental protocols, Biochim. Biophys. Acta 1367 (1998) 88–106.

48

ACCEPTED MANUSCRIPT [103] C. C. Moser, P. L. Dutton, Cytochrome c and c2 binding dynamics and electron transfer with photosynthetic reaction center protein and other integral membrane redox proteins,

PT

Biochemistry 27 (1988) 2450–2461.

RI

[104] L. Gerencsér, G. Laczkó, P. Maróti, (1999) Unbinding of oxidized cytochrome c from photosynthetic reaction center of Rhodobacter sphaeroides is the bottleneck of fast

SC

turnover, Biochemistry 38:16866–16875

NU

[105] J.W. Larson, C.A. Wraight, Preferential binding of equine ferricytochrome c to the bacterial photosynthetic reaction center from Rhodobacter sphaeroides, Biochemistry 39

MA

(2000) 14822–14830.

106] Z.S. Kolber, C.L. van Dover, R.A. Niederman, P.G. Falkowski, Bacterial photosynthesis in

TE

D

surface waters of the open ocean, Nature 407 (2000) 177– 179. [107] J.D. Olsen, J.D. Tucker, J.A. Timney, P. Qian, C. Vassilev, C.N. Hunter, The organization

AC CE P

of LH2 complexes in membranes from Rhodobacter sphaeroides, J. Biol. Chem. 283 (2008) 30772–30779.

[108] A.Verméglio, J. Lavergne, F. Rappaport, Connectivity of the intracytoplasmic membrane of Rhodobacter sphaeroides: a functional approach, Photosynth. Res. (2014) PMID: 25512104.

[109] S. Scheuring, R. Nevo, L.-N. Liu, S. Mangenot, D. Charuvi, T. Boudier V. Prima, P. Hubert, P, J.N. Sturgis, Z.Reich, The architecture of Rhodobacter sphaeroides chromatophores, Biochim Biophys Acta. 1837 (2014) 1263-1270. [110] R.C. Prince, A. Baccarini-Melandri, G.A. Hauska, B.A. Melandri, A.R Crofts, Asymmetry of an energy transducing membrane. The location of cytochrome c2 in Rhodopseudomonas spheroides and Rhodopseudomonas capsulata, Biochim Biophys Acta 387 (1975) 212-227.

49

ACCEPTED MANUSCRIPT [111] M.L. Cartron, J.D Olsen, M. Sener, P.J. Jackson, A.A. Brindley, P. Qian, M.J. Dickman, G.J., Leggett, K. Schulten, C.N. Hunter, Integration of energy and electron transfer

PT

processes in the photosynthetic membrane of Rhodobacter sphaeroides, Biochim. Biophys.

RI

Acta 1837 (2014) 1769-1780.

[112] L. Esser, M. Elberry, F. Zhou, C.-A. Yu, L. Yu, D. Xia,Inhibitor-complexed structures of

SC

the cytochrome bc1 from the photosynthetic bacterium Rhodobacter sphaeroides. J. Biol.

AC CE P

TE

D

MA

NU

Chem. 283 (2008) 2846–2857.

50

ACCEPTED MANUSCRIPT Figure Legends Fig. 1. High-resolution AFM topographs of the cytoplasmic surface of native ICM represented

PT

by Rba sphaeroides membrane patches (A-C) [14] and flattened thylakoid-like discs from Rps. photometricum (D-F) [15]. The Rba. sphaeroides membrane patches were prepared with 0.03%

SC

RI

-dodecyl maltoside, which is at a sub-critical micelle concentration, in order to open up ICM vesicles so that the resulting ICM patches are located closer to the mica surface, thereby

NU

maximizing the area available to the AFM tip during tapping mode acquisition of topographs. This represents a subsolubilizing level of detergent without apparent effects on membrane

MA

structure [10]. ICM preparations in panels (A) and (B) are from cells grown at low light intensity (4 W/m2). (C) Cells grown at high light intensity (220 W/m2). LH2 complexes are seen as

TE

D

separate nonameric ring structures (∼6–7 nm in diameter), assuming a zigzag appearance especially in LH2-only domains (B), as a result of vertical displacement when the natively

AC CE P

curved membranes adsorb onto the planar mica surface, leading to lines of raised LH2 complexes. This is also seen in panel (A) for a membrane patch also containing a row of RCLH1 dimers (∼8.5 nm RC-RC separation within dimer) with the RC-H subunit projecting from the membrane surface. Panel C shows a vast predominance of dimeric RC-LH1 complexes in patch form high light cells with only a few LH2 and monomeric RC-LH1 complexes. This predominance of RC-LH1 dimers was seen previously [16] in non-detergent treated membrane preparations representing nascent ICM growth initiation sites isolated from the upper-pigmented band. For the high- and low-light chromatophores, the LH2/LH1 molar ratios, determined as described by Sturgis et al. [17] were 0.53 and 1.23 respectively. (D-F) High-resolution AFM topographs of native Rsp. photometricum ICM preparations obtained by contact mode AFM. Here the RC-H subunit has been removed by nanodissection with the AFM tip leaving

51

ACCEPTED MANUSCRIPT cytoplasmic surfaces of the L and M subunits of the RC exposed. (D) High-light-adapted (~100 W/m2) ICM [15]. M, closed monomeric RC-LH1 core complex; LH2, nonameric LH2 complex.

PT

The core complexes and LH2 rings are arranged randomly and are found in an ∼1:3.5 ratio. (E,F)

RI

High-resolution topographs of low-light-adapted (10 to 20 W/m2) ICM. In panel E, an area of

SC

mixed core complexes and LH2 rings is shown, with an overall core complex:LH2 ratio of ∼1:7; areas of paracrystalline hexagonal arrays of LH2 are also seen that are devoid of core complexes.

NU

In panel F, an area is seen with substantially more core complexes than the average distribution. In these regions, as in panel D, several core complexes are found in contact, and the overall

MA

∼1:3.5 core complex:LH2 ratio is retained. The arrow in panel (D) delineates a free lipid bilayer space between LH2 complexes, which is thought to be functionally important in facilitating

TE

D

membrane mobility for quinone-quinol exchange to and from the core complex [18], just as the openings in the dimeric RC-LH1 complexes at the ends of linear core chains are thought to serve

AC CE P

in this function in Rba. sphaeroides [19].

Fig. 2. Proteomic analysis of chromatophore LH2 gel band derived from clear-native electrophoresis, showing the distribution of spectral counts [25]. Chromatophores were isolated after 3 and 24 h from cells undergoing acclimation from high to low light intensity. For clearnative electrophoresis, chromatophores were solubilized with n-octyl β-D-glucopyranoside (-OG) and deoxycholate (DOC), both at concentrations of 15 mM, and subjected to clear-native electrophoresis essentially as described by Wittig et al. [33]. Resolved gel bands were excised and subjected to in-gel trypsin digestion, followed by LC-MS/MS at sub-femtomol sensitivity. Tandem mass spectrometry data files were searched against the NCBI assembly of the annotated Rba. sphaeroides 2.4.1 genome [34]. Proteomic data were quantified by spectral counting, which

52

ACCEPTED MANUSCRIPT measures protein abundance on the basis of the number of tandem mass spectral observation for all constituent peptides. Accordingly, spectral counts mainly reflect the ability of trypsin to act at

PT

potential cleavage sites and are therefore only semiquantitative in nature. Proteins that are not

RI

highly abundant with a large number of accessible sites can give rise to higher counts than more abundant proteins with far fewer sites; in cases where no sites are available, no counts are found,

SC

e. g., PufB (LH1-β polypeptide) and PucA (LH2-α), although the latter polypeptide was detected

NU

after chymotrypsin cleavage [25]. Since the gel lanes were loaded with equal amounts of total protein, valid comparisons can be made between the same protein components in the

MA

chromatophore and upper-pigmented band preparations. Since spectral count sampling statistics have high technical reproducibility, they reflect precise assessments of relative protein

TE

D

abundance for differential protein expression studies such as those reported here. The spectral count distributions shown are for clusters of orthologous groups [35], in which the usual energy

AC CE P

production and conversion categories have been divided into subgroups, which account for the distinct metabolic capabilities unique to Rba. sphaeroides.

Fig. 3. Proteomic analysis of upper-pigmented (A) and chromatophore fractions (B) in the clear native electrophoresis RC-LH1 enriched gel band. Membrane fractions were isolated from cells after a 24-h adaptation to a light intensity of 100 W/m2 [32]; the respective membrane fractions were solubilized with digitonin and 

, and the resolved gel bands were subjected to in-

gel trypsin digestion and LC-MS/MS. The distributions shown are for clusters of orthologous groups [35], modified as described in the legend of Fig. 2.

Fig. 4. Effects of CCCP on the LH, RC and related energy transduction complexes of the ICM as

53

ACCEPTED MANUSCRIPT revealed by proteomic analysis of clear native electrophoresis gel bands resolved from the isolated chromatophore fraction. Chromatophores were solubilized with -OG/DOC and isolated

RI

PT

gel bands were subjected to in-gel trypsin digestion and LC-MS/MS [32].

SC

Fig. 5. Proteomic analysis of upper-pigmented fraction isolated from CCCP-treated cell suspensions undergoing ICM induction showing effect of CCCP on general membrane assembly

NU

factor levels (A,B). Following removal from sucrose gradients, the upper-pigmented band was subjected to sodium dodecyl sulfate-polyacrylamide gel electrophoresis and after entering, gel

MA

bands were excised for in-gel trypsin digestion followed by LC-MS/MS in which a 135-min LC

TE

D

gradient was used for optimal peptide resolution.

Fig. 6. (A, B) LH1-only membrane patches prepared using sub-critical micelle concentrations of

AC CE P

-dodecylmaltoglucoside, showing aberrant complexes as depicted in 3-D [53]. (A) LH-1 only membrane patch showing an incomplete LH1 ring (blue arrow), LH1 arcs (cyan arrows) and captive proteins (red arrows)(see also E, F). (B) LH-1 only membrane showing aberrant complexes consisting of a spiral structure and an open ring enclosed within the white and blue boxes, respectively. The noise levels were reduced in these topographs by using a low-pass filter. (C) Membrane patch from a LH1-limited mutant [53], showing a 3-D representation of an intact RC-LH1 complex (magenta arrow), an isolated RC (white arrow), an RC-LH1 complex without the H-subunit (blue arrow), and an empty LH1 ring (red arrow). (D-F) AFM topographs of LH1 complexes reconstituted in 2D LH1 crystals adsorbed onto mica, prepared from an LH1-only Rba. sphaeroides strain (Topographs kindly provided by Prof. C.N. Hunter). While the majority of complexes consisted of empty ()16 LH1 rings of ~115 Å diameter (D), the complexes

54

ACCEPTED MANUSCRIPT shown in E. and F. were among several that resemble putative assembly intermediates with a membrane protein entrapped in their interior, tentatively identified as LhaA, an LH1 assembly

PT

factor, which like the Lac permease belongs to the major facilitator superfamily, forming part of

RI

its putative ―Chl delivery branch‖ 52]. The space filling image (G) of the Lac permease [81], closely resembles the large transmembrane protein enclosed within the LH1 annulus (E, F),

SC

apparently representing the assembly factor still in association with its LH1 substrate. It is

NU

possible that entrapment of LhaA within the LH1 ring arose from the lack of RCs in this strain, which may normally displace this complex specific assembly factor during the core assembly

MA

process in which the RC is believed to act as a template [53].

TE

D

Fig. 7. Fluorescence kinetic transient elicited upon 450-nm excitation of Rba. sphaeroides cells undergoing induction of ICM development at low aeration [24] measured in a Satlantic

AC CE P

Fluorescence Induction/Relaxation system (Satlantic Inc., Halifax, Nova Scotia). Measurements were made on whole cells diluted typically to ~ 30 nM BChl in growth medium to produce transients in the linear range. A blue LED (450 nm, 30-nm bandwidth) served as the excitation source. Fluorescence emission is passed through an 880-nm interference filter (50-nm bandwidth) and detected by a sensitive avalanche photodiode module. The digitized fluorescence kinetic transients obtained at 880 nm are processed by computer-assisted analysis, which translates measured signals to programmed physiological parameters, following a charge separation elicited between the RC-BChl special pair and QA. Four phases can be distinguished in the plotted transients: phase I, induction phase (strong pulse of 144 μs duration), eliciting a single-turnover flash (STF), cumulatively saturating the photosystem to permit measurement of fluorescence induction from F0 to FM; phase II, relaxation phase in which indirect modulated

55

ACCEPTED MANUSCRIPT light is applied to assess the relaxation kinetics of the fluorescence yield on a 500 ms time scale, reflecting reopening of the RC; phase III, in which a 20 ms flash induces multiple turnovers

PT

saturating the photosystem and UQ pool (MTF, multiple turnover flash); and phase IV, applying

RI

indirect modulated light that records the kinetics of UQ pool reoxidation on a 1 s time scale. The transients represent signal-averaged sets of 20 traces per sample, which minimized noise levels

SC

of individual traces. Note the difference in the rate of fluorescence induction between the 0 and

NU

48 h samples, reflecting marked differences in the functional absorption cross-section. The registered variable fluorescence signal reflects the RC redox state in which a low fluorescence

MA

yield indicates an open RC (no charge, RC capable of performing photochemistry), while a high fluorescence yield indicates a closed RC (positively charged RC, transiently non-functional

TE

D

photochemistry). Parameters derived from the induction kinetics include: the quantum yield of the primary charge separation estimated from FV/FM where FV, the variable fluorescence, is the

AC CE P

difference between the minimal fluorescence (F0) and the maximal fluorescence (FM); the functional absorption cross section (), calculated from the slope of the single turnover saturation curve, related to the rate of increase in fluorescence yield. Since the extent to which the relaxation kinetics reflected RC-BChl special pair re-reduction and/or QA re-oxidation was previously unknown, the multiphasic relaxation profile is expressed as the RC electron transfer turnover rate, designated as ET. Analysis of the induction phase (I) gave functional absorption cross-section (450 nm) values of 40.6, 71.5 and 101.5 Å2 in cells sampled at 0, 3 and 48 h, respectively, while the respective values for the RC electron transfer turnover rate (ET) of 3.1, 1.7 and 6.0 ms were obtained from the relaxation phase (II). ET represents the average of the first and second phases of the three phases derived from the exponential decay kinetics.

56

ACCEPTED MANUSCRIPT Fig. 8. EM and AFM analyses of chromatophore membrane showing location of cytochrome bc1 complex labeled with gold NTA-Nanogold® [111]. (A) A negatively stained patch consisting of

PT

a single membrane bilayer with the gold-labeled cytochrome bc1 complexes mainly clustering

RI

near one edge. (B) The negatively stained features in A. were identified as the LH2 complex (green), RC–LH1 complex (blue/red), and cytochrome bc1 dimers (purple), identified from gold

SC

labeling. (C) Upper panel: close-up of false color 3-D depiction of AFM topograph obtained

NU

under liquid, showing intact, uncollapsed chromatophores labeled with nanogold. Image has been filtered with a low pass filter to reduce noise. A cluster of four putative nanogold beads

MA

indicating possible positions of dimeric bc1 complexes are contained within the red outline. The recessed area extending from the cluster shows a protein-free region. The pairs of colored

TE

D

asterisks denote the RC-H subunits of RC–LH1 core complexes; the 9-nm separation determined for the peaks indicated by the paired blue, red or magenta asterisks is consistent with the peak-to-

AC CE P

peak separation of H-subunits in a RC–LH1 dimer complex. Bottom panel: interpretation of AFM data in which putative nanogold beads are indicated by the orange circles denoting the possible positions of the cytochrome bc1 complexes contained within the red outline in the upper panel. The dimeric RC-LH1 complexes are also shown (blue/red). The edge-to-edge separation of the pairs of gold beads was found to be 2.4 ± 0.5 nm; compatible with the structure of the dimeric Rba. sphaeroides cytochrome bc1 complex [112].

57

Figure 1

AC CE P

TE

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

58

AC CE P

TE

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

Figure 2

59

AC CE P

TE

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

Figure 3

60

AC CE P

Figure 4

TE

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

61

AC CE P

TE

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

Figure 5

62

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

AC CE P

TE

Figure 6

63

AC CE P

Figure 7

TE

D

MA

NU

SC

RI

PT

ACCEPTED MANUSCRIPT

64

SC

RI

PT

ACCEPTED MANUSCRIPT

AC CE P

TE

D

MA

NU

Figure 8

65

PT

ACCEPTED MANUSCRIPT

SC

RI

Graphical Abstract

AC CE P

TE

D

MA

NU

Robert A. Niederman, Development and Dynamics of the Photosynthetic Apparatus in Purple Phototrophic Bacteria

66

ACCEPTED MANUSCRIPT

AC CE P

TE

D

MA

NU

SC

RI

PT

The author has no conflicts of interest associated with this study.

67