Skip to main content
UKPMC Funders Author Manuscripts logoLink to UKPMC Funders Author Manuscripts
. Author manuscript; available in PMC: 2025 Aug 13.
Published in final edited form as: Nature. 2019 Nov 13;575(7783):535–539. doi: 10.1038/s41586-019-1746-6

Cryo-EM structure of the spinach cytochrome b6 f complex at 3.6 Å resolution

Lorna A Malone 1, Pu Qian 1, Guy E Mayneord 1, Andrew Hitchcock 1, David A Farmer 1, Rebecca F Thompson 2, David JK Swainsbury 1, Neil A Ranson 2, C Neil Hunter 1,3,*, Matthew P Johnson 1,3,*
PMCID: PMC7617996  EMSID: EMS207426  PMID: 31723268

Abstract

The cytochrome b6 f (cytb6 f) complex has a central role in oxygenic photosynthesis, linking electron transfer between photosystems I and II and converting solar energy into a transmembrane proton gradient for ATP synthesis13. Electron transfer within cytb6 f occurs via the quinol (Q) cycle, which catalyses the oxidation of plastoquinol (PQH2) and the reduction of both plastocyanin (PC) and plastoquinone (PQ) at two separate sites via electron bifurcation2. In higher plants, cytb6 f also acts as a redox-sensing hub, pivotal to the regulation of light harvesting and cyclic electron transfer that protect against metabolic and environmental stresses3. Here we present a 3.6 Å resolution cryo-electron microscopy (cryo-EM) structure of the dimeric cytb6 f complex from spinach, which reveals the structural basis for operation of the Q cycle and its redox-sensing function. The complex contains up to three natively bound PQ molecules. The first, PQ1, is located in one cytb6 f monomer near the PQ oxidation site (Qp) adjacent to haem bp and chlorophyll a. Two conformations of the chlorophyll a phytyl tail were resolved, one that prevents access to the Qp site and another that permits it, supporting a gating function for the chlorophyll a involved in redox sensing. PQ2 straddles the intermonomer cavity, partially obstructing the PQ reduction site (Qn) on the PQ1 side and committing the electron transfer network to turnover at the occupied Qn site in the neighbouring monomer. A conformational switch involving the haem cn propionate promotes two-electron, two-proton reduction at the Qn site and avoids formation of the reactive intermediate semiquinone. The location of a tentatively assigned third PQ molecule is consistent with a transition between the Qp and Qn sites in opposite monomers during the Q cycle. The spinach cytb6 f structure therefore provides new insights into how the complex fulfils its catalytic and regulatory roles in photosynthesis.


Photosynthesis sustains life on Earth by converting light into chemical energy in the form of ATP and NADPH, producing oxygen as a by-product. Two light-powered electron transfer reactions at photosystems I and II (PSI and PSII) are linked via the cytb6 f complex to form the ‘Z-scheme’ of photosynthetic linear electron transfer (LET)1. Cytb6 f catalyses the rate-limiting step in LET, coupling the oxidation of PQH2 and reduction of PC and PQ to the generation of a transmembrane proton gradient, which is used by ATP synthase to make ATP2,3. The cytb6 f complex is analogous to the cytbc1 complex found in mitochondria4 and anoxygenic photosynthetic bacteria5, and both operate via the modified Q cycle2,6. The cytb6 f and cytbc1 complexes are dimeric and have similarly arranged electron transfer cofactors, comprising a 2Fe-2S cluster, two b-type haems and a c-type haem. However, crystallographic structures of cyanobacterial and algal cytb6 f complexes have revealed additional cofactors that are not found in cytbc1 complexes, including chlorophyll a, 9-cis β-carotene and an additional c-type high-spin haem79. The Q-cycle involves bifurcated transfer of two electrons, derived from oxidizing a lipophilic PQH2 molecule at the Qp-binding site, into the high- (2Fe-2S, cytf) and low- (cytbp, bn and cn) redox potential pathways, whereas the two protons enter the thylakoid lumen2,6. The high-potential pathway delivers an electron to a membrane-extrinsic soluble acceptor protein, PC, destined for PSI, while the low potential pathway delivers its electron to a PQ molecule bound at the Qn site near the stromal side of the membrane. Oxidation of a second PQH2 at the Qp site culminates in the two-electron reduction of a Qn site bound PQ, which together with two proton transfers from the stroma, regenerates PQH2. The Q-cycle thereby doubles the number of protons transferred to the lumen per PQH2 oxidized2,6. Yet, full understanding of the Q-cycle mechanism is hindered by a lack of information on the binding of the substrate PQ/ PQH2 molecules within the complex.

In addition to its role in LET, cytb6 f also plays a key part as a redox sensing hub involved in the regulation of light harvesting and cyclic electron transfer (CET), which optimize photosynthesis in fluctuating light environments10,11. Cytb6 f communicates the redox status of the PQ pool to the loosely associated light harvesting complex II (LHCII) kinase, STN71214. Phosphorylation of LHCII results in a decrease in thylakoid membrane stacking, promoting the exchange of LHCII between PSII and PSI to balance their relative excitation rates15 and regulate CET16. CET involves the reinjection of electrons from ferredoxin into the PQ pool, generating a proton gradient for photoprotective downregulation of PSI and PSII or to augment ATP synthesis, without net formation of NADPH11. The cytb6 f complex has been proposed to fulfil the role of the ferredoxin–PQ oxidoreductase (FQR) during CET, with the stromal-facing haem cn suggested to channel electrons from ferredoxin NADP+ reductase (FNR) bound ferredoxin to the Qn-site PQ17. How cytb6 f performs these central redox-sensing regulatory roles remains unclear.

Genetic manipulation of photosynthetic regulation is now recognized as being key to increasing crop yields to feed a global population projected to approach 10 billion by 205018. Indeed, overproduction of the Rieske iron–sulfur protein (ISP) of cytb6 f in Arabidopsis thaliana led to a 51% increase in yield19. Further progress in understanding the regulatory roles of cytb6 f and potentially manipulating them for crop improvement requires knowledge of the structure of the higher plant complex. Here, using a gentle purification procedure to obtain a highly active dimeric complex (Extended Data Fig. 1) and single-particle cryo-EM, we resolve the cytb6 f complex from Spinacia oleracea (spinach) at 3.6 Å resolution (Extended Data Fig. 2, Extended Data Table 1).

The colour-coded map (Fig. 1a–c) shows the architecture of this dimeric complex surrounded by a disordered density comprising detergent and lipid molecules. The overall organization of this higher plant cytb6 f complex is similar to crystallographic structures of algal and cyanobacterial complexes from Chlamydomonas reinhardtii7 (Protein Data Bank (PDB): 1Q90), Mastigocladus laminosus8 (PDB: 1VF5) and Nos-toc sp. PCC 71209 (PDB: 2ZT9) (Extended Data Table 2). Each monomeric unit of the cytb6 f complex comprises four large polypeptide subunits that contain redox co-factors (cytf, cytb6, ISP and subunit IV), and four small subunits (PetG, PetL, PetM and PetN). Extended Data Figure 3 shows the density and structural model for each subunit. The extrinsic domains of cytf and the ISP on the lumenal face of the complex flank the membrane-integral cytb6 subunits (Fig. 1a, b). The organization of the transmembrane integral subunits can be seen on the stromal side of the complex (Fig. 1c), with 13 transmembrane helices visible within each monomer (Fig. 1d–f). Peripheral to the core of cytb6 (four transmembrane helices) and subunit IV (three transmembrane helices) on the long axis of the complex is the single kinked transmembrane helix of the ISP that crosses over to provide the soluble ISP domain of the neighbouring monomer. The single transmembrane helix belonging to cytf is sandwiched by the transmembrane helices of the four minor subunits PetG, PetL, PetM and PetN, which form a ‘picket fence’ at the edge of the complex.

Fig. 1. Cryo-EM structure of the cytb6 f complex from spinach.

Fig. 1

ac, Views of the colour-coded cytb6 f density map showing cytb6 (green), cytf (magenta), ISP (yellow), subunit IV (cyan), PetG (grey), PetM (pink), PetN (pale orange) and PetL (pale purple). Detergent and other disordered molecules are shown in semi-transparent light grey. a, View in the plane of the membrane. The grey stripe indicates the probable position of the thylakoid membrane bilayer. b, View perpendicular to the membrane plane from the lumenal (p) side. c, View perpendicular to the membrane plane from the stromal (n) side. df, Modelled subunits of cytb6 f shown in a cartoon representation and coloured as in ac. d, View in the plane of the membrane. e, View perpendicular to the membrane plane from the lumenal side. f, View perpendicular to the membrane plane from the stromal side.

Figure 2a, b shows the organization of the prosthetic groups and lipids, with four c-type haems (f and cn, dark blue), four b-type haems (bp and bn, red), two 2Fe-2S clusters (burnt orange and yellow), two 9-cis β-carotenes (orange), two chlorophyll a molecules (green), three PQ molecules (yellow) and twelve bound lipids (two monogalactosyl diacylglycerol, four phosphatidylglycerol, three sulfoquinovosyl diacylglycerol and three phosphatidylcholine, all shown in white). Extended Data Figure 4 shows the density map and structural model for each prosthetic group. Figure 2c shows all of the bound electron transfer cofactor edge-to-edge distances within the cytb6 f complex. Electron transfer from the 2Fe-2S cluster is thought to involve movement of the lumenal ISP domain, pivoting between closer association with the Qp site and the haem f. In comparison to the chicken cytbc1 complex, in which the two conformations of the ISP were resolved20, the ISP and bound 2Fe-2S cluster in the spinach cytb6 f structure appear to be in the distal position with respect to haem f, as in the existing algal and cyanobacterial cytb6 f structures (Extended Data Table 2). PQ locations are generally inferred from crystallographic structures containing tightly bound quinone analogue inhibitors2123. The cryo-EM structure was obtained with native PQ molecules (Fig. 2d), clearly defined by their respective densities (Extended Data Fig. 4); their distances from the nearest cofactors are shown in Fig. 2e–g. One PQ molecule (PQ1) is adjacent to the haem bp and chlorophyll on one side of the dimer (Fig. 2e), and a second (PQ2) binds adjacent to the haem cn–haem bn pair on the opposite monomer to PQ1 (Fig. 2f). A third and less clearly defined PQ (PQ3) lies between the haem cn of one monomer and the haem bn of the other (Fig. 2g). The density map in this region could also be interpreted as a phospholipid; Extended Data Fig. 6 shows the two possible fits—to a plastoquinone or a lipid.

Fig. 2. The global arrangement of prosthetic groups, lipids and plastoquinone molecules in the spinach cytb6 f complex.

Fig. 2

a, b, The arrangement of molecules in the cytb6 f complex viewed in the membrane plane (a) and perpendicular to the membrane plane from the stromal side (b). Chl, chlorophyll a; bn, haem bn; cn, haem cn; bp, haem bp; β-car, 9-cis β-carotene; FeS, 2Fe-2S. c, d, Cofactors and edge-to-edge distances (in Å) in the dimeric cytb6 f complex. e, The location of the 1,4-benzoquinone ring of PQ1 adjacent to haem bp, the 2Fe-2S centre and two conformations of the chlorophyll molecule, represented in two shades of green. f, Close-up of the 1,4-benzoquinone ring of PQ2 and the nearby haem cn and haem bn near the stromal face of the complex. g, The 1,4-benzoquinone ring of PQ3, which sits between the haem cn and haem bn from the two cytb6 f monomers. The cytb6 f complex is coloured as in Fig. 1, and shows c-type haems (f and cn, dark blue), b-type haems (bp and bn, red), 9-cis-β-carotene (orange), chlorophyll a (green), 2Fe-2S (burnt orange and yellow), lipids (white) and plastoquinones (PQ1–PQ3; yellow).

The 1,4-benzoquinone ring of PQ1 is 16.2 Å from haem bp and 26.4 Å from the 2Fe-2S cluster (Fig. 2e), and distal to the Qp quinone oxidizing site defined in the M. laminosus cytb6 f structure23 (PDB: 4H13) by the inhibitor tridecylstigmatellin (Fig. 3a, b). The Qp site is located in a pocket formed by hydrophobic residues from subunit IV (Val84, Leu88, Val98 and Met101) and cytb6 (Phe81, Val126, Ala129, Val133, Val151 and Val154) (Fig. 3c). Bifurcated electron transfer to the 2Fe-2S cluster and haem bp involves two deprotonation events mediated by the ISP His128 and subunit IV Glu78 residues2,3, which are buried inside the Qp pocket (Fig. 3a, b). The OH group of PQ1 is ∼26 Å from His128, a ligand of the 2Fe-2S cluster (Fig. 3b), so PQ1 is unlikely to be oxidized in its resolved position, which probably represents a snapshot of its approach to the Qp site. It is notable in this regard that our spinach cytb6 f structure resolves two conformations of the chlorophyll phytyl tail, one of which permits access to Qp site and one that restricts it (Fig. 3c, d). There is only one position of the phytyl tail for the chlorophyll on the opposing monomer. The bound chlorophyll adjacent to PQ1 may fulfil a gating function at the Qp pocket, either controlling access of PQH2 and/or increasing the retention time of the reactive semiplastoquinone (SPQ) intermediate species formed following electron transfer to the 2Fe-2S cluster. Indeed, spin-coupling between the SPQ and the 2Fe-2S cluster has been detected during enzymatic turnover of cytb6 f but is absent in cytbc1 complexes that lack the chlorophyll molecule24. SPQ in the 2Fe-2S-bound state does not react with oxygen, providing a potential mechanism to control the release of superoxide from the Qp site24 and regulate the activity of the LHCII kinase STN725, which is proposed to bind to cytb6 f between transmembrane helices F and H of subunit IV26. Another role for chlorophyll in regulating the activity of STN7 could involve PQH2 displacing the chlorophyll phytyl tail on binding to the Qp site; this motion could induce a conformational change in STN7 leading to its activation27.

Fig. 3. Conformational alterations in the chlorophyll phytyl chain at the PQH2-oxidizing Qp site.

Fig. 3

a, Orientation of the PQ1 in relation to the haem bp, chlorophyll and 2Fe-2S cofactors. The catalytically essential residue E78 and coordinating residues of the 2Fe-2S cofactor are shown. Tridecylstigmatellin (TDS) is a quinone analogue, superimposed according to its position determined in the cyanobacterial complex (PDB: 4H13)23, and used here to indicate the probable destination of PQ1 in the Qp pocket. b, The same cofactors and residues as in a, but in relation to a surface view of cytb6 (green) and subunit IV (sub IV, cyan). c, The Qp pocket is highlighted with a red dashed line, showing its position in relation to the chlorophyll and PQ1 molecules; the hydrophobic residues of subunit IV (cyan) and cytb6 that line the pocket are shown as sticks and coloured cyan and green, respectively. d, The two conformations of the chlorophyll tail (represented in dark green and light green) gate (dashed arrow) the entrance to the Qp pocket.

PQ2 binds towards the stromal face of the complex, 4.4 Å from the haem cnbn pair at the Qn reducing site (Fig. 2f). The bn and cn haems on each monomer are separated by 4.9 Å, with the bn haem coordinated by His202 and His100 (cytb6), whereas the vinyl side-chain of haem cn is covalently linked to Cys35 (cytb6) (Fig. 2f). The dimerization interface of the cytb6 f complex creates a cavity, which is proposed to promote transfer of quinones between the Qp and Qn sites on neighbouring monomers8 (Fig. 4a, b). It is noteworthy that the three resolved PQ molecules inhabit this cavity and that PQ2 assumes a position ‘diagonally’ opposite to PQ1 (Fig. 4c) on the other monomer, as previously suggested28. PQ2 adopts a bowed conformation that straddles the intermonomer cavity with the distal PQ2 tail appearing to partially obstruct the Qn site in the neighbouring monomer (Fig. 4d, e). This arrangement may have functional importance in preventing the simultaneous binding of PQ molecules at both Qn sites, avoiding competition for electrons and favouring faster turnover of the Q cycle. Rapid provision of two electrons for PQ2 bound at a particular Qn site could be facilitated by the 15.3 Å electron-tunnelling distance between bp haems (Fig. 2c), which enables rapid inter-monomer electron transfer via the ‘bus-bar’ model from the neighbouring low-potential chain29,30. Alternatively, the second electron could be provided to the haem cn directly via an FNR–ferredoxin complex bound at the stromal surface via CET17,28. The haem cn propionates on the two halves of the cytb6 f dimer adopt different conformations (Fig. 4f, g); in the PQ-vacant site on the opposing monomer, the haem cn propionate is more closely associated with Arg207 (Fig. 4f), whereas in the PQ-occupied site, the haem cn propionate is rotated towards the 1,4-benzoquinone ring of PQ2 (Fig. 4g). The altered ligation of haem cn on PQ binding is consistent with the downshift of its redox potential31, which would strongly favour PQ reduction. We note that the reduction and oxidation of haem cn is accompanied by the binding and release of one proton31 so only one proton is required from the stromal side via the Arg207 and Asp20 residues (Fig. 4f, g) for PQ2 reduction to proceed rapidly, avoiding SPQ formation. It is also possible to position an oppositely oriented PQ within the density map, albeit with a less satisfactory fit (Extended Data Fig. 5). A third PQ molecule (PQ3) (Fig. 2g) has been assigned to the density between the Qp and Qn binding sites (see Extended Data Fig. 6 for an alternative assignment as phosphatidylcholine) with the 1,4-benzoquinone ring near the channel that links the two sides of the intermonomer cavity and the isoprenyl tail at the mouth of the Qp site. This third PQ may therefore capture a snapshot of the molecule transitioning between the Qp and Qn sites in opposite monomers.

Fig. 4. The intermonomer cavity of the spinach cytb6 f complex.

Fig. 4

a, b, Surface representations of the complex, with subunits coloured as in Fig. 1, and cofactors and lipids coloured as in Fig. 2. These two views of the complex are related by a 45° rotation about an axis perpendicular to the membrane to show two views of the cavity and the locations of PQ molecules. c, PQ1–PQ3 are shown in relation to the bn, cn and bp haems in the core of the complex, viewed in the membrane plane. d, The complex viewed from the stromal side of the membrane; peripheral helices of cytb6 and subunit IV are shown in cartoon representation for clarity, to show PQ2 straddling the intermonomer cavity and sitting between the two cn haems. e, Close-up of the cavity in d. f, g, The head and tail regions of PQ2 in relation to the cn haems on both sides of the cavity, highlighting the different orientations of the haem cn propionates, and the Arg207 and Asp20 side chains. The distances in angstroms between the residues and cofactors are labelled.

The cryo-EM structure of spinach cytb6 f reveals the positions of natively bound PQ and provides details regarding the conformational switches involved in PQ binding to the Qn site, chlorophyll gating of the Qp site and PQ exchange between the sites during Q-cycle operation.

Online content

Any methods, additional references, Nature Research reporting summaries, extended data, supplementary information, acknowledgements, peer review information; details of author contributions and competing interests; and statements of data and code availability are available at https://doi.org/10.1038/s41586-019-1746-6.

Methods

No statistical methods were used to predetermine sample size. The experiments were not randomized. The investigators were not blinded to allocation during experiments and outcome assessment.

Complex purification

Dimeric cytb6 f was isolated from dark-adapted market spinach (S. oleracea) in a procedure adapted from Dietrich and Kuhlbrandt32.

In brief, spinach leaves were homogenized in buffer 1 (50 mM Tris-HCl pH 7.5, 200 mM sucrose and 100 mM NaCl). Homogenate was then filtered and centrifuged for 15 min at 4,540g, 4 °C. Following centrifugation, the supernatant containing cell debris was discarded and the pellet resuspended in buffer 2 (10 mM Tricine-NaOH pH 8 and 150 mM NaCl) before centrifugation again for 15 min (4,540g, 4 °C). The resultant pellet was resuspended in buffer 3 (2 M NaBr, 10 mM Tricine-NaOH pH 8 and 300 mM sucrose) and incubated on ice for 15 min before diluting twofold with ice-cold milliQ H2O and centrifuging (15 min, 4,540g, 4 °C). The resultant pellet was resuspended in buffer 3 and incubated on ice for 15 min before diluting twofold with ice-cold milliQ H2O and centrifuging again (15 min, 4,540g, 4 °C). The pellet was resuspended in buffer 2 and centrifuged for 15 min, 4,540g at 4 °C. The final pellet was resuspended in a small volume of buffer 4 (40 mM Tricine pH 8.0, 10 mM MgCl2 and 10 mM KCl). The resultant thylakoid suspension was adjusted to 10 mg ml−1 chlorophyll (chlorophyll concentrations determined as described previously33).

For selective solubilization of cytb6 f, the thylakoid suspension (10 mg ml−1 chlorophyll) was diluted with membrane extraction buffer (40 mM Tricine pH 8.0, 10 mM MgCl2, 10 mM KCl and 1.25% (w/v) Hecameg) to a final concentration of 2 mg ml−1 chlorophyll, 1% (w/v) Hecameg. The resultant solution was mixed thoroughly then incubated for 2 min at room temperature before dilution to 0.75% (w/v) Hecameg with buffer 4. Unsolubilized material was removed by ultracentrifugation at 244,000g at 4 °C for 30 min in a Beckman Ti50.2 rotor.

The solubilization supernatant was concentrated using a Centriprep 100K centrifugal filter (Merck Millipore) before loading onto a 10–40% (w/v) continuous sucrose gradient containing 40 mM Tricine pH 8, 10 mM MgCl2, 10 mM KCl, 0.8% (w/v) Hecameg and 0.1 mg ml−1 egg yolk l-α-phosphatidylcholine (Sigma). This was ultracentrifuged at 174,587g at 4 °C for 16 h in a Beckman SW32 rotor.

A brown-ish band containing cytb6 f was collected from a region of the gradient corresponding to ∼16% sucrose. This band was concentrated and loaded onto a ceramic hydroxyapatite column (CHT) (Type I, Bio-Rad) pre-equilibrated in 20 mM Hecameg, 0.1 mg ml−1 phosphatidylcholine and 20 mM Tricine pH 8. The column was washed with 5 column volumes of CHT wash buffer (20 mM Hecameg, 0.1 mg ml−1 phosphatidylcholine and 100 mM ammonium phosphate pH 8) before bound material was eluted with CHT elution buffer (20 mM Hecameg, 0.1 mg ml−1 phosphatidylcholine and 400 mM ammonium phosphate pH 8).

Detergent exchange and gel filtration

Concentrated CHT eluate was loaded onto a 10–35% (w/v) continuous sucrose gradient containing 50 mM HEPES pH 8, 20 mM NaCl and 0.3 mM 4-trans-(4-trans-propylcyclohexyl)-cyclohexyl α-maltoside (tPCCαM) and ultracentrifuged at 175,117g at 4 °C for 16 h in a Beck-man SW41 rotor.

A single brown band containing cytb6 f was collected from a region of the gradient corresponding to ∼22% sucrose. This band was concentrated and loaded onto HiLoad 16/600 Superdex 200 pg gel filtration column (GE Healthcare) connected to an ÄKTA prime plus purification system (GE Healthcare). The column was run at a rate of 0.2 ml min−1 with 145 ml with gel filtration buffer (50 mM HEPES pH 8, 20 mM NaCl, 0.3 mM tPCCαM). Eluted fractions comprising dimeric cytb6 f were pooled and concentrated.

SDS–PAGE and BN-PAGE analysis

Samples collected from each purification step were analysed by SDS–PAGE and BN-PAGE. For SDS–PAGE, precast NuPAGE 12% Bis-Tris gels (Invitrogen) were run for 60 min at 180 V before staining with Coomassie blue. For BN-PAGE, precast NativePAGE 3–12% Bis-Tris gels (Invitrogen) were run for 120 min at 160 V before staining with Coomassie blue. Gels were imaged using an Amersham 600 imager (GE Healthcare).

Quantification of purified dimeric cytb6 f using redox difference spectra

Absorbance spectra were recorded at room temperature on a Cary60 spectrophotometer (Agilent). For redox difference spectra cytochromes were first fully oxidized with a few grains of potassium ferricyanide followed by reduction with a few grains of sodium ascorbate (cytf) then sodium dithionite (cytf and cytb6). At each stage the sample was mixed thoroughly and incubated for ∼1 min before recording spectra. Redox difference spectra (ascorbate-reduced minus ferricyanide-oxidized and dithionite-reduced minus ascorbate-reduced) were calculated and used to determine the concentrations of c-type haem of cytf and the two b-type haems of cytb6 using extinction coefficients of 25 mM cm−1 (haem f) and 21 mM cm−1 (cytb6 haems)34.

Reduction of decylplastoquinone

Approximately 0.1 mg decylplastoquinone (Merck) was dissolved in 100 μl ethanol, mixed with a few grains of sodium dithionite dissolved in 100 μl milliQ H2O and vortexed until the solution became colourless. Decylplastoquinol was extracted by mixing with 0.5 ml hexane, vortexing and centrifuging at 16,000g for 2 min. The hexane layer was carefully removed ensuring none of the aqueous phase was collected. Hexane extraction was repeated on the aqueous phase twice more, then the hexane solutions were pooled and dried in a rotary evaporator at 30 °C for 1 h before re-dissolving in ∼100 μl DMSO. Decylplastoquinol concentration was determined by diluting 10 μl of the DMSO solution into 795 μl ethanol, recording the absorbance spectrum between 250 and 350 nm and using an extinction coefficient35 of 3,540 M−1 cm−1 at 290 nm.

Purification of PC

PC was purified in its oxidized form from market spinach. In brief, spinach leaves were homogenized in buffer containing 50 mM sodium phosphate pH 7.4, 5 mM MgCl2 and 300 mM sucrose. Homogenate was then filtered and centrifuged for 15 min at 4,000g. Following centrifugation, the supernatant containing cell debris was discarded and the pellet was resuspended in buffer containing 10 mM Tricine pH 7.4 and 5 mM MgCl2. The solution was incubated on ice for 1 min before diluting twofold with buffer containing 10 mM Tricine pH 7.4, 5 mM MgCl2, 400 mM sucrose and centrifuging for 15 min at 4,000g. Following centrifugation, the pellet was resuspended to a chlorophyll concentration of 2 mg ml−1 in buffer containing 10 mM HEPES pH 7.6, 5 mM NaCl and 5 mM EDTA, and sonicated for 10 min, at 30 s intervals. The solution was centrifuged at 200,000g for 1 h to pellet any large unbroken material. The supernatant was applied to four 5-ml GE Healthcare Hi-TRAP Q FF anion-exchange columns in series, equilibrated in HEPES pH 8, 5 mM NaCl. A gradient of 0.005–1 M NaCl was used for elution, with PC eluting at around 200 mM. PC-containing fractions were identified by the blue colour on addition of potassium ferricyanide. These fractions were pooled, concentrated in a Vivaspin 3-kDa molecular-weight cut-off spin concentrator and loaded onto a Superdex 200 16/600 FPLC column, equilibrated with 20 mM HEPES pH 8 and 20 mM NaCl. The resulting PC fractions were pooled, concentrated and frozen at −80 °C until use.

Activity assays

Reduction of PC by cytb6 f was monitored by stopped-flow absorbance spectroscopy using an Olis RSM 1000 rapid-scanning spectrophotometer equipped with a USA-SF stopped flow cell at 20 °C. Solution A (231.25 nM cytb6 f and 62.5 μM PC in 50 mM HEPES pH 8, 20 mM NaCl and 0.3 mM tPCCαM) and solution B (1.25 mM decylplasto-quinol in the same buffer) were prepared and the reaction was initiated by mixing the solutions in a 4:1 volumetric ratio (final concentrations: 185 nM cytb6 f, 50 μM PC and 250 μM decylplastoquinol). PC reduction was monitored by recording absorbance spectra between 420 and 750 nm at a rate of 62 scans s−1 and plotting the change in absorbance36 at 597 nm. In a control reaction, cytb6 f was omitted to record the uncatalysed reduction of PC by decylplastoquinol. Fitting of the initial reaction rates was performed in Origin. All measurements were carried out in triplicate.

CryoEM specimen preparation and data acquisition

In brief, 3 μl of purified cytb6 f (∼17 μM) was applied to freshly glow-discharged holey carbon grids (Quantifoil R1.2/1.3, 400 mesh Cu). The grids were blotted for 2 s at 8 °C then plunge frozen into liquid ethane using a Leica EM GP at 90% relative humidity. Data acquisition was carried out on a Titan Krios microscope operated at 300 kV (Thermo Fisher) equipped with an energy filtered (slit width 20 eV) K2 summit direct electron detector. A total of 6,035 movies were collected in counting mode at a nominal magnification of 130,000× (pixel size of 1.065 Å) and a dose of 4.6 e Å−2 s−1 (see Extended Data Table 1). An exposure time of 12 s was used and the resulting movies were dose-fractionated into 48 fractions. A defocus range of −1.5 to −2.5 μm was used.

Image processing and 3D reconstruction

Beam-induced motion correction and dose-fractionation were carried out using MotionCor2. Contrast transfer function (CTF) parameters of the dose-weighted motion-corrected images were then estimated using GCTF37. All subsequent processing steps were performed using RELION 2.138 or 3.039 unless otherwise stated.

In total, 422,660 particles were manually picked from 6,035 micrographs. These particles were extracted using a box size of 220 × 220 pixels and subjected to reference-free 2D classification. A typical micro-graph showing picked particles is shown in Extended Data Fig. 2a, b. Particles that categorized into poorly defined classes were rejected, while the remaining 292,242 (69.2%) particles were used for further processing. A subset of 30,000 particles was used to generate a de novo initial model using the ‘3D initial model’ subroutine. The initial model low-pass filtered to 20 Å was used as a reference map for subsequent 3D classification into 10 3D classes. One stable 3D class at a resolution of 5.38 Å was selected for high-resolution 3D auto-refinement; this class accounted for a subset of 108,560 particles (25.6%). This subset of refined particles was then re-extracted and re-centred before another round of 3D auto-refinement was carried out. The resultant 4.85 Å density map was corrected for the modulation transfer function (MTF) of the Gatan K2 summit camera then further sharpened using the post-processing procedure to 4.02 Å. Per-particle CTF-refinement was carried out and a soft mask was created which included the detergent shell. The final global resolution estimate of 3.58 Å was based on the gold-standard Fourier shell correlation (FSC) cut-off of 0.143.

Local resolution was determined using one of two unfiltered half-maps as an input, a calibrated pixel size of 1.065 and a B-factor of −103. The output local resolution map is shown in Extended Data Fig. 2d, e.

Model building

Initially, a homology-based approach was performed using the crystal-lographic structure of Nostoc sp. PCC 7120 cytb6 f (PDB: 4OGQ)40 as a template. Sequence alignments of the eight polypeptide subunits of cytb6 f were carried out using Clustal Omega (Extended Data Figs. 7, 8). The model was rigid-body docked into the density using the ‘fit in map’ tool in Chimera41. This was then followed by manual adjustment and real-space refinement using COOT42. Sequence assignment and fitting was guided by bulky residues such as Arg, Trp, Tyr and Phe. After fitting of the polypeptide chains and cofactors in one half of the dimeric complex, the other half of the complex was then independently fitted into the C1 density map. Once both halves of the complex were fitted, cofactors, lipids and plastoquinone-9 molecules were fitted into regions of unassigned density. The final model underwent global refinement and minimization using the real space refinement tool in PHENIX43. The final refinement statistics are summarized in Extended Data Table 1.

Pigment analysis by reversed-phase HPLC

Pigments were extracted from purified cytb6 f with 7:2 acetone:methanol (v/v) and clarified extracts were separated by reversed-phase HPLC at a flow rate of 1 ml min−1 at 40 °C using a Supelco Discovery HS C18 column (5 μm particle size, 120 Å pore size, 25 cm × 4.6 mm) on an Agilent 1200 HPLC system. The column was equilibrated in acetonitrile: water:trimethylamine (9:1:0.01 v/v/v) and pigments were eluted by applying a linear gradient of 0–100% ethyl acetate over 15 min followed by isocratic elution with 100% ethyl acetate for a further 5 min. Elution of carotenoid and chlorophyll species was monitored by absorbance at 400, 450, 490 and 665 nm. Chlorophyll a was identified by its absorption spectra and known retention time44. The major carotenoid species was confirmed as 9-cis β-carotene using a standard obtained from Sigma-Aldrich (product no. 52824).

Extended Data

Extended Data Fig. 1. Purification of cytb6 f from spinach.

Extended Data Fig. 1

a, Absorption spectrum of ascorbate-reduced purified b6 f complex. The peak at 421 nm corresponds to the Soret band of bound pigments (chlorophyll a and haems). The peaks at 554 and 668 nm correspond to c-type haems of cytf and chlorophyll a, respectively. The inset panel shows redox difference spectra of ascorbate-reduced minus ferricyanide-oxidized b6 f (dashed line) and dithionite-reduced minus ascorbate-reduced (dotted line) cytb6 f. Redox difference spectra show haem f absorption peaks at 523 and 554 nm as well as absorption peaks at 534 and 563 nm corresponding to the b-type haems of cytb6. The calculated ratio of cytb6 b-type haems to the c-type haem of cytf was ∼2 using extinction coefficients of 25 mM cm−1 (f) and 21 mM cm−1 (b6)34. The spectra exhibit the absorption properties characteristic of intact cytb6 f. Spectra were recorded at room temperature. b, SDS–PAGE analysis of purified cytb6 f indicates that the sample is highly pure, with the four large subunits of the complex (cytf, cytb6, the Rieske ISP and subunit IV) running at ∼31 kDa, ∼24 kDa, ∼20 kDa and ∼17 kDa, respectively and the four small subunits (PetG, PetL, PetM and PetN) running at around 4 kDa (not shown). c, d, Negative-stain and BN-PAGE analysis of purified cytb6 f demonstrates the sample is dimeric and highly homogenous, with a single band corresponding to dimeric cytb6 f shown in lane 1. Lane 2 shows a sample that has been deliberately monomerized following incubation with 1% Triton X-100 for 1 h. For gel source data see Supplementary Fig. 1. e, The catalytic rate of plastocyanin reduction by the purified dimeric cytb6 f complex as determined by stopped-flow absorbance spectroscopy. A rate of 200 e s−1 was determined by taking the initial linear region from the enzyme-catalysed reaction (solid line) and subtracting the background rate measured in the absence of enzyme (long-dashed line). Plastocyanin reduction was not observed in the absence of decylplastoquinol (short-dashed line). Reactions were initiated upon addition of decylplastoquinol to the solution containing plastocyanin and b6 f while monitoring the loss of absorbance at 597 nm. Final concentrations were 50 μM plastocyanin, 185 nm b6 f and 250 μM decylplastoquinol. All experiments were performed in triplicate and controls were performed in the absence of b6 f or decylplastoquinol.

Extended Data Fig. 2. Cryo-EM micrographs of the spinach cytb6 f complex and calculation of the cryo-EM map global and local resolution.

Extended Data Fig. 2

a, Cytb6 f particles covered by a thin layer of vitreous ice on a supported carbon film. b, Examples of dimeric cytb6 f particles are circled in green. We recorded 6,035 cryo-EM movies, from which 422,660 particles were picked manually for reference-free 2D classification. The final density map was calculated from 108,560 particles. c, Gold-standard refinement was used for estimation of the final map resolution (solid black line). The global resolution of 3.58 Å was calculated using a FSC cut-off at 0.143. A model-to-map FSC curve (solid grey line) was also calculated. d, e, A C1 density map of the cytb6 f complex both with (d) and without (e) the detergent shell. The map is coloured according to local resolution estimated by RELION and viewed from within the plane of the membrane. The colour key on the right shows the local structural resolution in angstroms (Å).

Extended Data Fig. 3. Cryo-EM densities and structural models of polypeptides in the cytb6 f complex.

Extended Data Fig. 3

Polypeptides are coloured as in Fig. 1. The contour levels of the density maps were adjusted to 0.0144.

Extended Data Fig. 4. Cryo-EM densities and structural models of prosthetic groups, lipids and plastoquinone molecules in the cytb6 f complex.

Extended Data Fig. 4

c-type haems (f, cn; dark blue), b-type haems (bp, bn; red), 9-cis β-carotene (orange), chlorophyll a (major conformation, dark green; minor conformation, light green), 2Fe-2S (burnt orange and yellow), plastoquinones (yellow), monogalactosyl diacylglycerol (light pink), phosphatidylcholine (light cyan), sulfoquinovosyl diacylglycerol (light green) and phosphatidylglycerol (light purple). The contour levels of the density maps were adjusted to 0.0068.

Extended Data Fig. 5. Alternative interpretation of the region assigned as PQ2.

Extended Data Fig. 5

a, b, The density map showing two possible alternative conformations for PQ2, the major conformation (a) and the alternative conformation (b). Cofactors are coloured as in Extended Data Fig. 4 with b-type haems (bp and bn) coloured red, c-type haems (cn) coloured dark blue, chlorophyll a (major conformation) coloured dark green, plastoquinones coloured yellow and the cytb6 subunit coloured light green. The contour level of the density map was adjusted to 0.0089.

Extended Data Fig. 6. Alternative interpretations of the density map in the region assigned as PQ3.

Extended Data Fig. 6

a, b, The density map modelled with a plastoquinone molecule (a) and a phosphatidylcholine molecule (b). Top, the protein-free density map; bottom, the map including cytb6 (green). The 2.9 Å distance indicates a close contact between the PQ3 head group and the conserved Lys208. Cofactors are coloured as in Extended Data Fig. 4 with b-type haems (bp and bn) coloured red, chlorophyll a (major conformation) in dark green, plastoquinones in yellow, phosphatidylcholine in light cyan, sulfoquinovosyl diacylglycerol in mint green and the cytb6 subunit in light green. The contour level of the density map was adjusted to 0.0127.

Extended Data Fig. 7. Multiple sequence alignment of cytb6 f subunits cytf and cytb6.

Extended Data Fig. 7

a, b, Sequences of cytf (a) and cytb6 (b) from cyanobacterial (M. laminosus and Nostoc sp. PCC7120), algal (C. reinhardtii) and plant (S. oleracea) subunits were aligned in Clustal Omega v.1.2.4. Conserved identities are indicated by asterisks, and similarities by double or single dots. Polar residues are coloured in green, positively charged residues are pink, hydrophobic residues are red and negatively charged residues are blue. The sequences omit signal peptides.

Extended Data Fig. 8. Multiple sequence alignment of the Rieske ISP, subunit IV, PetG, PetL, PetM and PetN.

Extended Data Fig. 8

af, Sequences of Rieske ISP (a), subunit IV (b), PetG (c), PetL (d), PetM (e) and PetN (f) from cyanobacterial (M. laminosus and Nostoc sp. PCC7120), algal (C. reinhardtii) and plant (S. oleracea) subunits were aligned in Clustal Omega v.1.2.4. Conserved identities are indicated by asterisks, and similarities by double or single dots. Polar residues are coloured in green, positively charged residues are pink, hydrophobic residues are red and negatively charged residues are blue. The sequences omit signal peptides.

Extended Data Table 1. Cryo-EM data collection, refinement and validation statistics.

S. oleracea cytb6f
(EMD-4981)
(PDB 6RQF)
Data collection and processing
Magnification 130,000 X
Voltage (kV) 300
Electron exposure (e2) 1.15 (55.2 e- on 48 frames)
Defocus range (μm) -1.5 to-2.5
Pixel size (Å) 1.065
Symmetry imposed Cl
Initial particle images (no.) 422,660
Final particle images (no.) 108,560
Map resolution (Å) 3.58
  FSC threshold 0.143
Map resolution range (Å) ~3.3-8.3
Refinement
Initial model used (PDB code) RELION de novo model from 30,000 particles
Model resolution (Å) 3.58
  FSC threshold 0.143
Model resolution range (Å) ~3.3-8.7
Map sharpening B factor (Å2) Estimated automatically using RELION*
Model composition
  Non-hydrogen atoms 16,359
  Protein residues 1,944
  Ligands 29
B factors (Å2)
  Protein RELION auto-estimated
  Ligand RELION auto-estimated
R.m.s. deviations (PHENIX)
  Bond lengths (Å) 0.009
  Bond angles (°) 1.182
Validation
  MolProbity score 1.83
  Clashscore 9.67
  Poor rotamers (%) 0.18
Ramachandran plot
  Favored (%) 95.40
  Allowed (%) 4.24
  Disallowed (%) 0.37
*

Data from ref. 45.

Extended Data Table 2. Comparison of cofactor distances in b6f and bc1 dimers from different species.

(PDB 6RQF) (PDB 1Q90) (PDB 2E74) (PDB 4OGQ)
Source S. oleracea C. reinhardtii M. laminosas Nostoc sp. PCC 7120
Resolution (Å) 3.6 3.1 3.0 2.5
Inhibitors * - TDS (Qp) - -
Distances:
bn- cn (Å) 4.7, 4.7 4.7, 4.7 4.7, 4.7 4.6, 4.6
bn- bp (Å) 12.1, 12.0 12.2, 12.2 12.2, 12.2 12.1, 12.1
bp- bp (Å) 15.3 15.1 15.2 15.3
bp - [2Fe-2S] (Å) 25.6, 25.5 22.9, 22.9 25.5, 25.5 25.3, 25.3
[2Fe-2S]-f (Å) 25.9, 26.1 27.8, 27.8 26.2, 26.2 26.2, 26.2
(PDB1BCC) (distal) (PDB 3BCC) (proximal)
Source G. gallus G. gallus
Resolution (Å) 3.2 3.7
Inhibitors * - STG (Qp), AMY (Qn)
Distances:
bn- bp (Å) 12.4, 12.4 12.3, 12.3
bp- bp (Å) 14.4 14.5
bp- [2Fe-2S] (Å) 30.3, 30.3 23.0, 23.1
[2Fe-2S]-c1 (Å) 16.8, 16.8 27.3, 27.5

The distances are edge-to-edge (Å), for each half of the b6f dimer from different species (PDB: 6RQF, 1Q90, 2E74, 4OGQ) and the bc1 dimer from Gallus gallus with the Rieske ISP in its distal (PDB: 1BCC) and proximal (PDB: 3BCC) positions. AMY, antimycin; STG, stigmatellin.

Supplementary Material

Supplementary Material

Reporting summary.

Further information on research design is available in the Nature Research Reporting Summary linked to this paper.

Acknowledgements

M.P.J. acknowledges funding from the Leverhulme Trust grant RPG-2016-161. C.N.H., P.Q., A.H., D.J.K.S. and M.P.J. also acknowledge financial support from the Biotechnology and Biological Sciences Research Council (BBSRC UK) award numbers BB/M000265/1 and BB/P002005/1. L.A.M. was supported by a White Rose doctoral studentship, G.E.M. was supported by a doctoral studentship from The Grantham Foundation and D.A.F. was supported by a University of Sheffield doctoral scholarship. Cryo-EM data was collected at the Astbury Biostructure Laboratory funded by the University of Leeds (ABSL award) and the Wellcome Trust (108466/Z/15/Z). We thank S. Tzokov, J. Bergeron, J. Wilson and D. Mann for their helpful advice and assistance with the EM and data processing.

Footnotes

Author contributions P.Q., C.N.H., N.A.R. and M.P.J. supervised the project. L.A.M., G.E.M., P.Q., C.N.H., R.F.T. and M.P.J. designed the experiments. L.A.M. and G.E.M. purified the cytb6 f complex, L.A.M., G.E.M., A.H. and D.J.K.S. characterized the cytb6 f complex. L.A.M., P.Q., D.A.F. and R.F.T. collected, processed and/or analysed the cryo-EM data. L.A.M., C.N.H. and M.P.J. wrote the manuscript. All authors proofread and approved the manuscript.

Competing interests The authors declare no competing interests.

Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Peer review information Nature thanks Zhenfeng Liu, Alexander Tikhonov and the other, anonymous, reviewer(s) for their contribution to the peer review of this work.

Reprints and permissions information is available at https://www.nature.com/nature-portfolio/reprints-and-permissions.

Data availability

All relevant data are available from the authors and/or are included with the manuscript or in the Supplementary Information. Atomic coordinates and the cryo-EM density map have been deposited in the Protein Data Bank under accession number 6RQF and the Electron Microscopy Data Bank (EMDB) under accession number EMD-4981.

References

  • 1.Hill R, Bendall F. Function of the two cytochrome components in chloroplasts: a working hypothesis. Nature. 1960;186:136–137. [Google Scholar]
  • 2.Cramer WA, Hasan SS, Yamashita E. The Q cycle of cytochrome bc complexes: a structure perspective. Biochim Biophys Acta. 2011;1807:788–802. doi: 10.1016/j.bbabio.2011.02.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 3.Tikhonov AN. The cytochrome b6f complex at the crossroad of photosynthetic electron transport pathways. Plant Physiol Biochem. 2014;81:163–183. doi: 10.1016/j.plaphy.2013.12.011. [DOI] [PubMed] [Google Scholar]
  • 4.Xia D, et al. Crystal structure of the cytochrome bc1 complex from bovine heart mitochondria. Science. 1997;277:60–66. doi: 10.1126/science.277.5322.60. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Esser L, et al. Inhibitor-complexed structures of the cytochrome bc1 from the photosynthetic bacterium Rhodobacter sphaeroides. J Biol Chem. 2008;283:2846–2857. doi: 10.1074/jbc.M708608200. [DOI] [PubMed] [Google Scholar]
  • 6.Cape JL, Bowman MK, Kramer DM. Understanding the cytochrome bc complexes by what they don’t do. The Q-cycle at 30 Trends Plant Sci. 2006;11:46–55. doi: 10.1016/j.tplants.2005.11.007. [DOI] [PubMed] [Google Scholar]
  • 7.Stroebel D, Choquet Y, Popot JL, Picot D. An atypical haem in the cytochrome b6f complex. Nature. 2003;426:413–418. doi: 10.1038/nature02155. [DOI] [PubMed] [Google Scholar]
  • 8.Kurisu G, Zhang H, Smith JL, Cramer WA. Structure of the cytochrome b6f complex of oxygenic photosynthesis: tuning the cavity. Science. 2003;302:1009–1014. doi: 10.1126/science.1090165. [DOI] [PubMed] [Google Scholar]
  • 9.Baniulis D, et al. Structure–function, stability, and chemical modification of the cyanobacterial cytochrome b6f complex from Nostoc sp. PCC 7120. J Biol Chem. 2009;284:9861–9869. doi: 10.1074/jbc.M809196200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 10.Bellafiore S, Barneche F, Peltier G, Rochaix JD. State transitions and light adaptation require chloroplast thylakoid protein kinase STN7. Nature. 2005;433:892–895. doi: 10.1038/nature03286. [DOI] [PubMed] [Google Scholar]
  • 11.Yamori W, Shikanai T. Physiological functions of cyclic electron transport around photosystem I in sustaining photosynthesis and plant growth. Annu Rev Plant Biol. 2016;67:81–106. doi: 10.1146/annurev-arplant-043015-112002. [DOI] [PubMed] [Google Scholar]
  • 12.Horton P, Black MT. Activation of adenosine 5′-triphosphate induced quenching of chlorophyll fluorescence by reduced plastoquinone. The basis of state I–state II transitions in chloroplasts. FEBS Lett. 1980;119:141–144. [Google Scholar]
  • 13.Vener AV, van Kan PJM, Rich PR, Ohad I, Andersson B. Plastoquinol at the quinol oxidation site of reduced cytochrome bf mediates signal transduction between light and protein phosphorylation: thylakoid protein kinase deactivation by a single-turnover flash. Proc Natl Acad Sci USA. 1997;94:1585–1590. doi: 10.1073/pnas.94.4.1585. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 14.Gal A, Hauska G, Herrmann R, Ohad I. Interaction between light harvesting chlorophyll-a/b protein (LHCII) kinase and cytochrome b6/f complex. In vitro control of kinase activity. J Biol Chem. 1990;265:19742–19749. [PubMed] [Google Scholar]
  • 15.Allen JF, Bennett J, Steinback KE, Arntzen CJ. Chloroplast protein phosphorylation couples plastoquinone redox state to distribution of excitation energy between photosystems. Nature. 1981;291:25–29. [Google Scholar]
  • 16.Wood WHJ, et al. Dynamic thylakoid stacking regulates the balance between linear and cyclic photosynthetic electron transfer. Nat Plants. 2018;4:116–127. doi: 10.1038/s41477-017-0092-7. [DOI] [PubMed] [Google Scholar]
  • 17.Zhang H, Whitelegge JP, Cramer WA. Ferredoxin:NADP+ oxidoreductase is a subunit of the chloroplast cytochrome b6f complex. J Biol Chem. 2001;276:38159–38165. doi: 10.1074/jbc.M105454200. [DOI] [PubMed] [Google Scholar]
  • 18.Zhu XG, Long SP, Ort DR. Improving photosynthetic efficiency for greater yield. Annu Rev Plant Biol. 2010;61:235–261. doi: 10.1146/annurev-arplant-042809-112206. [DOI] [PubMed] [Google Scholar]
  • 19.Simkin AJ, McAusland L, Lawson T, Raines CA. Overexpression of the RieskeFeS protein increases electron transport rates and biomass yield. Plant Physiol. 2017;175:134–145. doi: 10.1104/pp.17.00622. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Zhang Z, et al. Electron transfer by domain movement in cytochrome bc1. Nature. 1998;392:677–684. doi: 10.1038/33612. [DOI] [PubMed] [Google Scholar]
  • 21.Yan J, Kurisu G, Cramer WA. Intraprotein transfer of the quinone analogue inhibitor 2,5-dibromo-3-methyl-6-isopropyl-p-benzoquinone in the cytochrome b6 f complex. Proc Natl Acad Sci USA. 2006;103:69–74. doi: 10.1073/pnas.0504909102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 22.Yamashita E, Zhang H, Cramer WA. Structure of the cytochrome b6f complex: quinone analogue inhibitors as ligands of heme cn. J Mol Biol. 2007;370:39–52. doi: 10.1016/j.jmb.2007.04.011. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Hasan SS, Yamashita E, Baniulis D, Cramer WA. Quinone-dependent proton transfer pathways in the photosynthetic cytochrome b6 f complex. Proc Natl Acad Sci USA. 2013;110:4297–4302. doi: 10.1073/pnas.1222248110. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Sarewicz M, et al. Metastable radical state, nonreactive with oxygen, is inherent to catalysis by respiratory and photosynthetic cytochromes bc1/b6f. Proc Natl Acad Sci USA. 2017;114:1323–1328. doi: 10.1073/pnas.1618840114. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25.Singh SK, et al. Trans-membrane signalling in photosynthetic state transitions: redox- and structure-dependent interaction in vitro between Stt7 kinase and the cytochrome b6f complex. J Biol Chem. 2016;291:21740–21750. doi: 10.1074/jbc.M116.732545. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 26.Zito F, Vinh J, Popot JL, Finazzi G. Chimeric fusions of subunit IV and PetL in the b6f complex of Chlamydomonas reinhardtii: structural implications and consequences on state transitions. J Biol Chem. 2002;277:12446–12455. doi: 10.1074/jbc.M110914200. [DOI] [PubMed] [Google Scholar]
  • 27.Saif Hasan S, Yamashita E, Cramer WA. Transmembrane signaling and assembly of the cytochrome b6f-lipidic charge transfer complex. Biochim Biophys Acta. 2013;1827:1295–1308. doi: 10.1016/j.bbabio.2013.03.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28.Nawrocki WJ, et al. The mechanism of cyclic electron flow. Biochim Biophys Acta. 2019;1860:433–438. doi: 10.1016/j.bbabio.2018.12.005. [DOI] [PubMed] [Google Scholar]
  • 29.Osyczka A, Moser CC, Daldal F, Dutton PL. Reversible redox energy coupling in electron transfer chains. Nature. 2004;427:607–612. doi: 10.1038/nature02242. [DOI] [PubMed] [Google Scholar]
  • 30.Świerczek M, et al. An electronic bus bar lies in the core of cytochrome bc1. Science. 2010;329:451–454. doi: 10.1126/science.1190899. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Alric J, Pierre Y, Picot D, Lavergne J, Rappaport F. Spectral and redox characterization of the heme ci of the cytochrome b6f complex. Proc Natl Acad Sci USA. 2005;102:15860–15865. doi: 10.1073/pnas.0508102102. [DOI] [PMC free article] [PubMed] [Google Scholar]

References

  • 32.Dietrich J, Kühlbrandt W. Purification and two-dimensional crystallization of highly active cytochrome b6 f complex from spinach. FEBS Lett. 1999;463:97–102. doi: 10.1016/s0014-5793(99)01609-9. [DOI] [PubMed] [Google Scholar]
  • 33.Porra RJ, Thompson WA, Kriedemann PE. Determination of accurate extinction coefficients and simultaneous equations for assaying chlorophylls a and b extracted with four different solvents: verification of the concentration of chlorophyll standards by atomic absorption spectroscopy. Biochim Biophys Acta. 1989;975:384–394. [Google Scholar]
  • 34.Cramer WA, Whitmarsh J. Photosynthetic cytochromes. Annu Rev Plant Physiol. 1977;28:133–172. [Google Scholar]
  • 35.Dawson RMC, Elliot DC, Elliot WH, Jones KM. Data for Biochemical Research. Clarendon; 1986. [Google Scholar]
  • 36.Tan S, Ho KK. Purification of an acidic plastocyanin from Microcystis aeruginosa. Biochim Biophys Acta. 1989;973:111–117. doi: 10.1016/s0005-2728(89)80410-4. [DOI] [PubMed] [Google Scholar]
  • 37.Zhang K. Gctf: Real-time CTF determination and correction. J Struct Biol. 2016;193:1–12. doi: 10.1016/j.jsb.2015.11.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Fernandez-Leiro R, Scheres SHW. A pipeline approach to single-particle processing in RELION. Acta Crystallogr D. 2017;73:496–502. doi: 10.1107/S2059798316019276. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 39.Zivanov J, et al. New tools for automated high-resolution cryo-EM structure determination in RELION-3. eLife. 2018;7:e42166. doi: 10.7554/eLife.42166. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Hasan SS, Cramer WA. Internal lipid architecture of the hetero-oligomeric cytochrome b6 f complex. Structure. 2014;22:1008–1015. doi: 10.1016/j.str.2014.05.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 41.Pettersen EF, et al. UCSF Chimera—a visualization system for exploratory research and analysis. J Comput Chem. 2004;25:1605–1612. doi: 10.1002/jcc.20084. [DOI] [PubMed] [Google Scholar]
  • 42.Adams PD, et al. PHENIX: a comprehensive Python-based system for macromolecular structure solution. Acta Crystallogr D. 2010;66:213–221. doi: 10.1107/S0907444909052925. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 43.Emsley P, Cowtan K. Coot: model-building tools for molecular graphics. Acta Crystallogr D. 2004;60:2126–2132. doi: 10.1107/S0907444904019158. [DOI] [PubMed] [Google Scholar]
  • 44.Proctor MS, et al. Plant and algal chlorophyll synthases function in Synechocystis and interact with the YidC/Alb3 membrane insertase. FEBS Lett. 2018;592:3062–3073. doi: 10.1002/1873-3468.13222. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.Rosenthal PB, Henderson R. Optimal determination of particle orientation, absolute hand and contrast loss in single particle electron cryomicroscopy. J Mol Biol. 2003;333:721–745. doi: 10.1016/j.jmb.2003.07.013. [DOI] [PubMed] [Google Scholar]

Associated Data

This section collects any data citations, data availability statements, or supplementary materials included in this article.

Supplementary Materials

Supplementary Material

Data Availability Statement

All relevant data are available from the authors and/or are included with the manuscript or in the Supplementary Information. Atomic coordinates and the cryo-EM density map have been deposited in the Protein Data Bank under accession number 6RQF and the Electron Microscopy Data Bank (EMDB) under accession number EMD-4981.

RESOURCES