You are on page 1of 24

Article

Structures of fungal and plant


acetohydroxyacid synthases

https://doi.org/10.1038/s41586-020-2514-3 Thierry Lonhienne1,7 ✉, Yu Shang Low1,7, Mario D. Garcia1, Tristan Croll2, Yan Gao3, Quan Wang3,
Lou Brillault4, Craig M. Williams1, James A. Fraser1, Ross P. McGeary1, Nicholas P. West1,
Received: 12 September 2019
Michael J. Landsberg1, Zihe Rao3,5,6, Gerhard Schenk1 & Luke W. Guddat1 ✉
Accepted: 1 June 2020

Published online: xx xx xxxx


Acetohydroxyacid synthase (AHAS), also known as acetolactate synthase, is a flavin
Check for updates adenine dinucleotide-, thiamine diphosphate- and magnesium-dependent enzyme
that catalyses the first step in the biosynthesis of branched-chain amino acids1. It is the
target for more than 50 commercial herbicides2. AHAS requires both catalytic and
regulatory subunits for maximal activity and functionality. Here we describe
structures of the hexadecameric AHAS complexes of Saccharomyces cerevisiae and
dodecameric AHAS complexes of Arabidopsis thaliana. We found that the regulatory
subunits of these AHAS complexes form a core to which the catalytic subunit dimers
are attached, adopting the shape of a Maltese cross. The structures show how the
catalytic and regulatory subunits communicate with each other to provide a pathway
for activation and for feedback inhibition by branched-chain amino acids. We also
show that the AHAS complex of Mycobacterium tuberculosis adopts a similar
structure, thus demonstrating that the overall AHAS architecture is conserved across
kingdoms.

AHAS (enzyme classification code, EC 2.2.1.6) is the first enzyme in and CSUs of Escherichia coli could combine16. However, the structure
the biosynthesis pathway for branched-chain amino acids (BCAAs)1 of the CSU–RSU complex has not yet been determined.
that leads to the production of l-valine, l-leucine and l-isoleucine. It Here, the crystal structure of ScAHAS (~800 kDa) and the cryo-electron
is present in plants and microorganisms, but not animals. Owing to its microscopy (cryo-EM) structures of AtAHAS (~710 kDa) in the presence
importance in plants, AHAS has been used as a target for 58 commercial or absence of l-valine have been determined (Extended Data Tables 1,
herbicides2, and is an emerging target for antimicrobial agents3. How- 2). In addition, we determined a crystal structure of a fungal hybrid
ever, with concerns over the spread of resistance to such herbicides4, it AHAS complex that consists of the CSU of Cryptococcus neoformans
is important to improve our understanding of the functional features (CnAHAS) and the ScRSU in the presence of l-valine (Extended Data
of this enzyme. Table 1). The ScAHAS complex was purified (Supplementary Figs. 3a,
AHAS consists of two subunits: a catalytic subunit (CSU) that 4a) in 50 mM phosphate (Supplementary Fig. 3b) and crystals were
contains flavin adenine dinucleotide (FAD), thiamine diphosphate obtained in the presence of ATP and the herbicide bensulfuron methyl
(ThDP) and magnesium (Mg2+)5,6, and a regulatory subunit (RSU) (BSM)2,17. The ScAHAS structure has eight CSUs and eight RSUs arranged
that enhances the activity of the CSU1,7,8. The RSUs also contain the in a Maltese cross shape, in which the four CSU dimers are attached at
binding sites for BCAAs, which are feedback inhibitors 1. Notably, the the extremities and the eight RSUs form the central core to which pairs
mechanism by which the RSUs exert their function(s) has remained of ATP are bound (Fig. 1a). A similar subunit arrangement occurs in
unclear. AtAHAS (Fig. 1b, c), except that the eight RSUs of ScAHAS are replaced
The CSUs consist of three domains—α, β and γ—that have highly con- by four AtAHAS RSUs, each of which contains two pseudo-repeats of the
served amino acid sequences and polypeptide lengths (Supplementary ScAHAS sequence (Supplementary Fig. 2). Furthermore, negative-stain
Fig. 1). By comparison, the RSUs are variable (Supplementary Fig. 2), microscopy illustrates that the AHAS complex of Mycobacterium tuber-
with sizes ranging from 11 to 60 kDa1. RSUs have ACT domain(s)9 that culosis has a similar shape (Supplementary Fig. 5), thus demonstrating
contain the BCAA-binding site9–11. In plant AHASs, there are two ACT that this overall arrangement of subunits is common across kingdoms.
domains in a single polypeptide, whereas microbial AHASs have only
one (Supplementary Fig. 2).
Structures of the CSUs from the AHASs of Arabidopsis thaliana The hexadecameric ScAHAS complex
(AtAHAS)2,12, Saccharomyces cerevisiae (ScAHAS)13 and Candida In the ScAHAS complex, N- and C-terminal extensions of the RSUs,
albicans3, and the RSUs from four bacterial AHASs10,14,15 have been deter- the ACT domains within the RSUs and the ATP-binding site provide
mined. In addition, a model has been suggested to show how the RSUs major contributions to the assembly of the complex. Each RSU interacts

School of Chemistry and Molecular Biosciences, The University of Queensland, Brisbane, Queensland, Australia. 2Cambridge Institute for Medical Research, University of Cambridge, Cambridge,
1

UK. 3Shanghai Institute for Advanced Immunochemical Studies, School of Life Science and Technology, ShanghaiTech University, Shanghai, China. 4Centre for Microscopy and Microanalysis, The
University of Queensland, Brisbane, Queensland, Australia. 5State Key Laboratory of Medicinal Chemical Biology, College of Life Science, Nankai University, Tianjin, China. 6Laboratory of Structural
Biology, Tsinghua University, Beijing, China. 7These authors contributed equally: Thierry Lonhienne, Yu Shang Low ✉e-mail: t.lonhienne@uq.edu.au; luke.guddat@uq.edu.au

Nature | www.nature.com | 1
Article
a ScCSU dimer b AtCSU dimer that interacts with several RSUs (see ‘Interactions between CSUs and
ScRSU AtRSU RSUs’), thus stabilizing the complex.
octamer tetramer

Conformation of the RSUs in the ScAHAS complex


ScCSU dimer

ScCSU dimer

AtCSU dimer

AtCSU dimer
The RSUs in ScAHAS have four distinct regions: an N-terminal extension
and domain, and a C-terminal extension and domain (Supplementary
Fig. 6a). The N-terminal and C-terminal domains have similar topolo-
ATP ×2 gies to those observed in the bacterial RSU homodimers14,15, and have
an N-terminal domain formed by two α-helices and four β-strands that
AtCSU dimer are arranged in the βαββαβ arrangement (Supplementary Fig. 6b) that
ScCSU dimer
is characteristic of ACT domains9,18. The conformation of the C-terminal
c AtCSU dimer domain shares a high level of structural similarity with its counterpart
AtRSU in bacteria, except that there is an extra sequence inserted in the ScRSU
tetramer that forms a protruding ‘contact’ loop (Supplementary Fig. 6c). This
loop interacts with the α–β linker of the CSU and with the N-terminal
extension of neighbouring RSUs.
90°
AtCSU dimer

AtCSU dimer

195 Å
Interactions between CSUs and RSUs
The main region of the RSUs that interacts with the CSUs is the ACT
domain (Extended Data Fig. 3a and Supplementary Fig. 7), which
is consistent with a previously proposed model for an E. coli AHAS
complex16. Notably, the first α-helix of the ACT domain (α1; residues
90–101) forms a salt bridge (R101 (RSU)–D199 (CSU)) with the ‘active
site’ Q-loop (191–204)19 (Extended Data Fig. 3b). F201 and Q202, located
AtCSU dimer
76 Å at the top of the Q-loop, are essential for enzymatic activity20. Q202 has
an important role in the stabilization of the substrate and/or intermedi-
Fig. 1 | The structure of the ScAHAS and AtAHAS complexes. a, ScAHAS. Eight
ate within the active site19,20, and F201 has a role in the stabilization of
RSUs in different colours—each of which contains a single ACT domain—form a
central core to which four CSU dimers are attached. Four ATP pairs (red
FAD and the substrate20,21. The pairing of R101 (RSU) and D199 (CSU) is
transparent surface) are bound to the RSU dimers. b, AtAHAS. Four RSUs highly conserved in AHAS complexes (Supplementary Figs. 1, 2), which
(magenta and green)—each of which contains two pseudo-repeats of the ACT emphasizes the importance of this interaction. In addition, the second
domain—form the central core. c. The cryo-EM map of AtAHAS. α-helix of the ACT domain (α2, residues 129–140 (RSU)) interacts with
the β-domain of one of the two CSUs, principally using N134 (CSU)
and R137 (CSU) (Extended Data Fig. 3b). Furthermore, two β-strands
with six RSUs and three CSUs (Supplementary Table 1). Both the N of the ACT domain, the contact loop and the N-terminal extension—all
and C termini of the RSUs adopt tentacle-like structures that extend of which are from different RSUs—form a cluster of residues that inter-
beyond the core regions of the RSUs and establish interactions with act with and stabilize the α–β linker of the CSUs (Extended Data Fig. 3c).
neighbouring subunits (Extended Data Fig. 1a). Four pairs of ATP also
contribute to the association between RSUs (Figs. 1a, 2). The RSUs
alone form dimers, which assemble into an octameric conformation Activation of the ScAHAS by the RSUs
that is dependent on the presence of ATP (Extended Data Fig. 1b). The Two distinct factors contribute to the enhancement of ScAHAS activity:
ATPs form pairs linked by a Mg2+ ion (Fig. 2a and Extended Data Fig. 1c, the dimerization of CSUs and allosteric communication within a dimer.
d), while the adenine and ribosyl moieties associate with all four RSUs First, dimerization is essential to form fully functional active sites13,
to complete a network of interactions (Fig. 2a and Extended Data and dimer stabilization is strongly promoted by the presence of RSUs
Fig. 1d). There are five key residues in ScAHAS that interact with ATP. and ATP (their presence decreases the dissociation constant (Kd) of the
Of these, three (R159, F232 and R280) are invariant (Supplementary CSUs by approximately 20-fold; Extended Data Fig. 4a). Second, the
Fig. 2), and one (R254) is highly conserved. Neither ATP nor the RSUs conformation of the CSU dimer ensures that the movements in the two
alone can enhance AHAS activity (Fig. 2b). However, in combination active sites are complementary and synchronized13,20,22; when the first
ATP and the RSUs cooperatively assemble and activate the AHAS step (decarboxylation of pyruvate) occurs in one active site, the second
complex (Fig. 2b). step (carbon–carbon bond formation) proceeds in the other in a con-
certed manner20. Complex formation triggers an approximately fourfold
increase in activity that is largely entropy-driven (Extended Data Fig. 4b, c)
Comparison of CSU structures with or without RSUs and is mediated by a bridge formed by the RSUs that directly connects
The fold of the CSUs, the conformation of the active site, and the loca- the two active sites of the CSU dimer (Fig. 3). This involves an interaction
tion of ThDP, Mg2+ and FAD are unaltered in the presence or absence between an α-helix (α1) from the two ACT domains and the Q-loops of
of RSUs17 (Extended Data Fig. 2a). Accordingly, the Michaelis–Menten the CSUs (that is, the R101 (RSU)–D199 (CSU) salt bridge; Fig. 3b). The
constant (KM) for the substrate (pyruvate) is similar in the presence importance of this salt bridge for the regulation of the CSU by the RSUs
and absence of the RSUs (3.2 and 3.0 mM, respectively; Extended Data was confirmed by assessing the interactions between the CSU and the
Fig. 2b). However, a substantial structural change occurs in one of the extended ScRSU form with an R101A mutation (ScRSUE(R101A)). Kinetic
regions that involves the linker between the α- and β-domains (α–β assays found two marked effects: the affinity between ScRSUE(R101A)
linker) of the CSUs13 (residues 258–279) when the RSUs bind. In the and CSU is around 20 times weaker than the affinity of the native com-
structure of the CSU alone, or in complex with BSM, this linker is ran- plex (Extended Data Fig. 4d) and the R101A mutation decreases the
dom coil, in which the six central residues are completely disordered. catalytic rate constant (kcat) twofold (Extended Data Fig. 4e). These
In the complex, this region has a helical fold (Extended Data Fig. 2c) results suggest that the role of the α1-bridge (Fig. 3b) is important to

2 | Nature | www.nature.com
a
Chain H

F232
Chain G ATP
R254
K251
R159

Mg2+ R280
R280

ATP
R254 K251

R159

Chain D
F232

Chain C
b
25 Kd = 1.25 ±0.02 μM
25 Kd = 36.7 ± 0.6 μM
Specific activity (U mg–1)

Specific activity (U mg–1)

20
20
ATP 1 mM ATP
15 15
ADP No ATP

10 10

5 5

0 0
0 20 40 60 80 100 120 140 0 1 2 3 4 5
ATP/ADP (μM) ScRSU (μM)

Fig. 2 | The role of ATP in assembling the ScAHAS complex. a, Binding mode cooperatively to the RSU (Hill coefficient = 3.0 ± 0.1; Kd = 36.7 ± 0.6 μM). ADP
of ATP. Four RSUs (chains C, D, G and H) interact with an ATP pair coordinated has no effect on AHAS activity. Right, the specific activity of ScAHAS (0.13 μM
by Mg 2+. The electron density (2Fo − Fc map) is shown in grey. b, Activation of in the assay) was measured with increasing RSU concentrations and a
ScAHAS by ATP and the RSU. The specific activity of AHAS was measured at saturating concentration of ATP (1 mM). RSU binding is cooperative (Hill
30 °C by monitoring pyruvate consumption. Left, the specific activity of coefficient = 3.1 ± 0.1; Kd = 1.25 ± 0.02 μM). The RSU in the absence of ATP had no
ScAHAS (0.13 μM in the assay) was measured with increasing ATP effect on activity. Data are mean ± s.e.m. of technical triplicates. Data were
concentrations and a saturating RSU concentration (8.5 μM). ATP binds fitted to the Hill equation to obtain the Kd and the Hill coefficient.

increase the turnover of the enzyme, presumably by improving the com- map (Fig. 1c and Supplementary Fig. 9) revealed that most of the poly-
munication between the active sites of CSU dimers. Other regions of the peptide in the two subunits is visible (Fig. 1b), except for the region that
RSUs (that is, the other α-helix (α2) and two of the β-strands of the ACT corresponds to the linker between the RSU repeats (Extended Data
domain) interact with the α- and β-domains of a CSU dimer (Extended Table 2). The overall arrangement of the RSUs and CSUs is similar to
Data Fig. 3b, c). These interactions rigidify regions of the dimer the ScAHAS complex, except that there is no contact loop in AtAHAS
that are not directly involved in the communication between the (Supplementary Fig. 10). Most features that have roles in the activation
active sites. Therefore, the apparent effect of the RSU is to decrease of ScAHAS are also preserved in the AtAHAS complex. In comparison
the entropy of the enzyme–substrate complex before activation, ena- with the CSU alone (Protein Data Bank (PDB) code 5K6Q), the complex
bling a more-efficient channelling of the energy that is transmitted from shows that a contraction of the CSU dimer occurs when the RSU binds
one active site to the other. The structural and kinetic data show that (Supplementary Fig. 11). Thus, the RSU stabilizes the CSU in AtAHAS
the increase in AHAS activity after the formation of the AHAS complex complex in the same manner as proposed for the ScAHAS complex.
is due to contacts that restrict the population of unfavoured conforma- Analogous to ScAHAS (Fig. 3), R110 (RSU repeat one) and R343 (RSU
tional states for catalysis and enhance the allosteric communication23 repeat two) each form a salt bridge with D204 (Q-loop) from opposing
between catalytic centres. We therefore proposed that the primary role CSUs (Supplementary Fig. 12). The weak affinity of the R110A and R343A
of the RSUs is to enhance the efficiency of the allosteric communication double mutant for the CSU prevented the isolation of the complex
between active sites (Extended Data Fig. 5). (Supplementary Fig. 13a). This confirms, as in the ScAHAS complex
(Extended Data Fig. 4d), that the R–D salt bridge is essential for the
stabilization of the complex.
The dodecameric AtAHAS complex The most notable difference between the ScAHAS and AtAHAS com-
The cryo-EM structure of the AtAHAS complex was obtained without plexes is the presence and absence of ATP, respectively. However, visual
herbicide (BSM) or ATP (Fig. 4b and Supplementary Fig. 8). The cryo-EM inspection shows that there is room to accommodate an ATP pair in the

Nature | www.nature.com | 3
Article
a a Q-loops
ThDP ThDP
L-Valine (site 1)
FAD FAD
D204 D204
R343 R110
α1 α1
α2 α2
90°

CSU α- and ACT domains


CSU α- and L-Valine (site 2)
β-domain β-domain
stabilization
stabilization b
V485 V485
ThDP ThDP
b C2 C2
FAD 7.15
5.07 FAD 7.15
5.07
He-ThDP ThDP 5.96 5.52 5.96 5.52
4.39 6.56 4.39 6.56
G121 G121
FAD FAD F206 F206

Q207 Q207

D204 D204

N217 N217
D199 D199 Fig. 4 | The effects of l-valine binding on the AtAHAS complex.
a, Superimposition of the ACT domains of the free AtAHAS complex and the
Q-loop complex with bound l-valine. The RSUs are shown in dark green and magenta
R101 R93 R93 R101 Q-loop
for the repeats in the structure of the complex in the absence of l-valine, and
cyan and coral for the repeats in the structure of the complex with bound
l-valine. l-Valine (ball and stick representation) is shown with a light-yellow
α1 α1 transparent surface around the outside. The Q-loops of the CSU dimer are
shown as stick models. The binding of l-valine triggers the disruption of the
α1-bridge
R110/R343–D204 salt bridge and the bending (black arrows) of the α-helices in
Fig. 3 | The ScRSU influences the conformation of the ScCSU. a, The two CSU the ACT domains. b, Stereoimage of the superimposed active sites. The
dimer polypeptides are shown in dark grey and light brown. The γ subunit is displayed residues are involved in substrate binding (green and light blue for
omitted for clarity. The two chains of the RSU dimer are shown in green and the complexes in the absence and presence of l-valine, respectively). In the
magenta. The two α 1 helices of the ACT domains in the RSUs establish a direct structure of the complex without l-valine, FAD and ThDP (ball and stick models)
path between the two catalytic Q-loops of the CSU dimer. The remaining are shown in yellow and coral, respectively. In the structure of the complex with
regions of the ACT domains are involved in the stabilization of the α- and l-valine, FAD and ThDP have similar but lighter-colour shades. The image shows
β-domains of the CSU dimer (enclosed in dashed grey ellipses). b, Communication that the active site expands after the binding of l-valine. C2 denotes the
paths between the Q-loops of the CSU dimer in the ScAHAS complex. The carbanion of the ylid that nucleophilically attacks the carbonyl of the substrate.
structure of ScAHAS (CSU alone) in complex with pyruvate (shown in green All distances are stated in angstroms.
ball-and-stick format) that represents the enzyme state during catalysis
(PDB code 5INU)20 is superimposed on a CSU dimer of the ScAHAS (CSU + RSU)
complex. The model displays the asymmetry that occurs between catalytic dimer in the free AtAHAS complex (Fig. 4a and Extended Data Fig. 7a).
centres, each of which has a different conformation to accomplish the two Another important feature is that the bending of the α1 helix—which is
half-cycles of the reaction at the same time. The representation of the CSU transmitted to the α2 helix (Fig. 4a and Extended Data Fig. 7b)—induces
dimer is restricted to the Q-loops and polypeptide extensions on both sides. a reorganization of the interactions between the ACT domain of the
The polypeptide extensions from both Q-loops interact with each other to RSUs and the β-domain of the CSUs (Supplementary Fig. 16). This
form the path through which the active sites communicate. Thus, the α 1-bridge
causes distancing (around 3 Å) between the β-domain and ACT domain
of the RSUs binds to the Q-loops (that is, R101–D199 and R93–N217) and
(Supplementary Fig. 16a) and leads to the expansion of the CSU dimer
enhances the communication between active sites.
(Extended Data Fig. 6c and Supplementary Fig. 11). A destabilization
of the CSUs is suggested as there is a decrease in the interface area
between CSU subunits (by 7%) and between the RSU and CSU dimer
AtAHAS complex (Supplementary Fig. 14). Thus, the precise role of ATP (by 14%) (Supplementary Fig. 17). These changes probably account for
in the AtAHAS complex should be investigated further. the partial dissociation of the CSU from the complex in the presence
of 5 mM l-valine (Supplementary Fig. 13b). Furthermore, there is an
The AtAHAS complex with bound l-valine expansion of the active sites as the distances between the reactive
A cryo-EM map of the AtAHAS complex obtained in the presence of carbon of ThDP and key residues that are involved in the binding of
l-valine (Extended Data Table 2 and Supplementary Fig. 15) shows the substrate20 increase by around 1.5 Å (Fig. 4b).
two densities in adjacent ACT domains where l-valine could be fit- Taken together, these structural observations provide several mecha-
ted (Extended Data Fig. 6a, b). The effect of l-valine is to induce bend- nisms by which feedback inhibition by BCAAs occurs. First, the altera-
ing of the α1 helices (Fig. 4a and Extended Data Fig. 6c). This triggers tion of the communication between the active sites, coupled with the
distancing (around 5 Å) between R110 (RSU) and D204 (CSU), which destabilization of the CSU dimer, is expected to reduce kcat. This hypoth-
causes the disruption of the α1-bridge that links the Q-loops of the CSU esis is supported by data obtained for the ScAHAS complex, in which

4 | Nature | www.nature.com
the increase in the stability of the CSU dimer (Extended Data Fig. 4b, c) acknowledgements, peer review information; details of author con-
and the connection between active sites made by the α1-bridge of the tributions and competing interests; and statements of data and code
RSUs (Extended Data Fig. 4e) are central features that are involved availability are available at https://doi.org/10.1038/s41586-020-2514-3.
in the upregulation of kcat. Second, the deformation of the active site
after the addition of l-valine (2 mM) leads to an increase in the KM for 1. Duggleby, R. G. & Pang, S. S. Acetohydroxyacid synthase. J. Biochem. Mol. Biol. 33, 1–36
(2000).
pyruvate (from 12.6 ± 1.9 to 24.2 ± 3.9 mM; Supplementary Fig. 18),
2. Garcia, M. D., Nouwens, A., Lonhienne, T. G. & Guddat, L. W. Comprehensive
and potentially also decreases the catalytic rate. Third, the weakening understanding of acetohydroxyacid synthase inhibition by different herbicide families.
of the interaction between the CSU dimer and the RSU promotes the Proc. Natl Acad. Sci. USA 114, E1091–E1100 (2017).
3. Garcia, M. D. et al. Commercial AHAS-inhibiting herbicides are promising drug leads for
dissociation of the CSUs from the complex. Although only a partial
the treatment of human fungal pathogenic infections. Proc. Natl Acad. Sci. USA 115,
dissociation is observed in the presence of 5 mM l-valine, it is possible E9649–E9658 (2018).
that this dissociation is more complete under physiological conditions. 4. Nandula, V. K. et al. Herbicide metabolism: crop selectivity, bioactivation, weed
resistance, and regulation. Weed Sci. 67, 149–175 (2019).
These interpretations are corroborated by a crystal structure
5. Schloss, J. V., Ciskanik, L. M. & Van Dyk, D. E. Origin of the herbicide binding site of
obtained for a fungal hybrid AHAS complex that consists of the CnCSU acetolactate synthase. Nature 331, 360–362 (1988).
and ScRSU in the presence of l-valine (Extended Data Fig. 8a and 6. Lonhienne, T., Garcia, M. D. & Guddat, L. W. The role of a FAD cofactor in the regulation of
acetohydroxyacid synthase by redox signaling molecules. J. Biol. Chem. 292, 5101–5109
Extended Data Table 1). Although it is a hybrid from two different spe-
(2017).
cies, kinetic studies show that like the ScAHAS complex, this complex 7. Pang, S. S. & Duggleby, R. G. Regulation of yeast acetohydroxyacid synthase by valine
is activated by ATP and inhibited by l-valine (Extended Data Fig. 8b). and ATP. Biochem. J. 357, 749–757 (2001).
8. Lee, Y. T. & Duggleby, R. G. Identification of the regulatory subunit of Arabidopsis thaliana
Density for l-valine is observed and its location is similar to that in acetohydroxyacid synthase and reconstitution with its catalytic subunit. Biochemistry 40,
the AtAHAS complex (Extended Data Fig. 8c). Notably, the structural 6836–6844 (2001).
changes observed after l-valine binding correlate with the observa- 9. Lang, E. J. M., Cross, P. J., Mittelstädt, G., Jameson, G. B. & Parker, E. J. Allosteric ACTion:
the varied ACT domains regulating enzymes of amino-acid metabolism. Curr. Opin.
tions observed for AtAHAS (Extended Data Fig. 8d). This includes the Struct. Biol. 29, 102–111 (2014).
disruption of the α1-bridge, the presence of which is essential for the 10. Bansal, A., Karanth, N. M., Demeler, B., Schindelin, H. & Sarma, S. P. Crystallographic
regulatory mechanism. structures of IlvN·Val/Ile complexes: conformational selectivity for feedback inhibition of
aceto hydroxyl acid synthases. Biochemistry 58, 1992–2008 (2019).
Previously, a minimal subunit arrangement for the AHAS complex 11. Aravind, L. & Koonin, E. V. Gleaning non-trivial structural, functional and evolutionary
had been proposed that involved only two to four RSUs and CSUs8,16. information about proteins by iterative database searches. J. Mol. Biol. 287, 1023–1040
Here we demonstrate that AHAS adopts a higher-order structure that (1999).
12. McCourt, J. A. & Duggleby, R. G. Acetohydroxyacid synthase and its role in the
is conserved across three different kingdoms. The formation of a RSU biosynthetic pathway for branched-chain amino acids. Amino Acids 31, 173–210 (2006).
multimeric core platform is proposed to have two advantages. First, 13. Lonhienne, T., Garcia, M. D., Fraser, J. A., Williams, C. M. & Guddat, L. W. The 2.0 Å X-ray
it contributes to the stabilization of both the ACT domains and CSUs. structure for yeast acetohydroxyacid synthase provides new insights into its cofactor and
quaternary structure requirements. PLoS ONE 12, e0171443 (2017).
In ScAHAS, this is achieved by the N- and C-terminal extensions that 14. Kaplun, A. et al. Structure of the regulatory subunit of acetohydroxyacid synthase
interact with distal ACT domains (Extended Data Fig. 1a). The linker isozyme III from Escherichia coli. J. Mol. Biol. 357, 951–963 (2006).
between repeats in AtAHAS may have a similar role. The stabilization 15. Petkowski, J. J. et al. Crystal structures of TM0549 and NE1324—two orthologs of E. coli
AHAS isozyme III small regulatory subunit. Protein Sci. 16, 1360–1367 (2007).
of the CSUs is a major factor for the regulation of activity. The second 16. Xie, Y. et al. Interactions between the ACT domains and catalytic subunits of
advantage of the RSU core structure—at least in ScAHAS—is that it ena- acetohydroxyacid synthases (AHASs) from different species. ChemBioChem 19,
bles a higher level of regulation through the creation of a binding site 2387–2394 (2018).
17. Lonhienne, T. et al. Commercial herbicides can trigger the oxidative inactivation of
for ATP. The BCAA biosynthetic pathway is an anabolic process that acetohydroxyacid synthase. Angew. Chem. Int. Ed. 55, 4247–4251 (2016).
requires energy, and it is therefore logical that downregulation occurs 18. Barak, Z. & Chipman, D. M. Allosteric regulation in acetohydroxyacid synthases
in response to low levels of energy (ATP) in the cell. The formation of (AHASs)—different structures and kinetic behavior in isozymes in the same organisms.
Arch. Biochem. Biophys. 519, 167–174 (2012).
the core formed by four RSU dimers has created an opportunity for 19. Lonhienne, T. et al. Structural insights into the mechanism of inhibition of AHAS by
a pocket to accommodate two paired ATPs, suggesting a previously herbicides. Proc. Natl Acad. Sci. USA 115, E1945–E1954 (2018).
unknown function for ATP. Overall, the active and regulatory sites 20. Lonhienne, T. et al. High resolution crystal structures of the acetohydroxyacid
synthase-pyruvate complex provide new insights into its catalytic mechanism.
(that is, Q-loop and ACT domain) in AHAS are shaped and positioned ChemistrySelect 2, 11981–11988 (2017).
for functional simplicity, with only minimal movements required for 21. Pang, S. S., Duggleby, R. G. & Guddat, L. W. Crystal structure of yeast acetohydroxyacid
their interaction. By contrast, when the feedback inhibitor—l-valine— synthase: a target for herbicidal inhibitors. J. Mol. Biol. 317, 249–262 (2002).
22. Belenky, I. et al. Many of the functional differences between acetohydroxyacid synthase
binds to the RSUs it triggers considerable structural changes, which (AHAS) isozyme I and other AHASs are a result of the rapid formation and breakdown of
result in the perturbation of interactions between the CSUs and lead the covalent acetolactate–thiamin diphosphate adduct in AHAS I. FEBS J. 279, 1967–1979
(2012).
to a reduction in catalytic activity.
23. Dai, S. et al. Low-barrier hydrogen bonds in enzyme cooperativity. Nature 573, 609–613
(2019).

Online content Publisher’s note Springer Nature remains neutral with regard to jurisdictional claims in
published maps and institutional affiliations.
Any methods, additional references, Nature Research reporting sum-
maries, source data, extended data, supplementary information, © The Author(s), under exclusive licence to Springer Nature Limited 2020

Nature | www.nature.com | 5
Article
Australian Government Department of Agriculture and Water Resources (CT-06). We thank the
Bio-Electron Microscopy facility (iHuman Institute) at ShanghaiTech University and the
Reporting summary
Cryo-EM Facility Center of Southern University of Science and Technology (Shenzhen) for
Further information on research design is available in the Nature providing support and acknowledge the facilities and scientific and technical assistance of the
Research Reporting Summary linked to this paper. Australian Microscopy & Microanalysis Research Facility at the Centre for Microscopy and
Microanalysis, The University of Queensland. T.C. is supported by a Wellcome Trust grant
(209407/Z/17/Z), awarded to R. Read. We thank R. Duggleby for initiating the early functional
and structural studies on AHASs.
Data availability
The structures of the ScAHAS complex, CnCSU–ScRSU–valine complex Author contributions L.W.G., T.L., G.S., Z.R., C.M.W., R.P.M., J.A.F. and N.P.W. conceived the
project. T.L. and Y.S.L. established the protein-purification protocols, generated mutants and
and AtAHAS in the presence and absence of l-valine have been depos- performed kinetic experiments. M.D.G. performed the initial purification and cryo-EM
ited in the Protein Data Bank with accession numbers 6U9D and 6WO1, experiments with AtAHAS. Y.S.L., Y.G., Q.W. and M.J.L. collected and processed the cryo-EM
6VZ8 and 6U9H, respectively. The cryo-EM map has been deposited in data. T.L. crystallized and solved the structures of the fungal AHASs. T.L. and L.W.G. collected
the X-ray data for the fungal AHASs. T.L., Y.S.L., L.W.G., G.S. and M.J.L. interpreted the data. L.B.
the Electron Microscopy Data Bank under accession numbers EMDB- assisted with cryo-EM and ReFyn analysis. T.C. performed the final refinement steps for the
21487 and EMDB-20700. Plasmids, strains and raw data are available cryo-EM structure of AtAHAS in complex with l-valine. T.L. and Y.S.L. prepared the figures. T.L.,
Y.S.L., L.W.G. and G.S. wrote the paper and T.L., Y.S.L., L.W.G., G.S., C.M.W., M.J.L., Z.R., N.P.W.,
from the corresponding author upon reasonable request.
J.A.F., R.P.M. and T.C. edited the paper.

24. Lonhienne, T., Gerday, C. & Feller, G. Psychrophilic enzymes: revisiting the
thermodynamic parameters of activation may explain local flexibility. Biochim. Biophys. Competing interests The authors declare no competing interests.
Acta 1543, 1–10 (2000).
Additional information
Supplementary information is available for this paper at https://doi.org/10.1038/s41586-020-
Acknowledgements Crystallization was performed at the University of Queensland 2514-3.
Remote-Operation Crystallization and X-ray diffraction facility (UQROCX). Data collection was Correspondence and requests for materials should be addressed to T.L. or L.W.G.
carried out at the Australian Synchrotron, part of ANSTO and made use of the Australian Peer review information Nature thanks Robert Sammons, Kai Tittmann and the other,
Cancer Research Foundation (ACRF) detector. This work was supported by funds from the anonymous, reviewer(s) for their contribution to the peer review of this work.
National Health and Medical Research Council (grant numbers 1087713 and 1147297), and the Reprints and permissions information is available at http://www.nature.com/reprints.
Extended Data Fig. 1 | See next page for caption.
Article
Extended Data Fig. 1 | The assembly of the ScAHAS complex. a, General view In the presence of 1 mM ATP (right), there are three peaks that correspond to
of the interactions that stabilize the ScAHAS complex. The 16 chains in the the dimer (34%), the tetramer (13%) and the octamer (49%). Partial assembly of
complex are signified by the letters A–P. Each RSU has N-terminal and the octamer was observed because the RSU could only be used at a maximum
C-terminal extensions that span through the core of the structure. As an of 130 nM in this experiment due to saturation of the detector (Supplementary
example, two regions of the N-terminal extension of the RSU (chain L) form Methods). c, d, Stereo views of the binding of ATP to the ScAHAS complex.
interactions with four neighbouring subunits (chains D (top inset), O, P (middle c, The coordination of the six phosphates by Mg 2+ and the stabilization of their
inset) and M (bottom inset)). b, The assembly of the ScRSU octamer monitored negative charges by six arginine and two lysine residues from two different
by mass photometry (ReFeyn, see Supplementary Methods). For this RSUs. d, Stabilization of the adenine and ribosyl moieties of one of the ATPs by
experiment, the extended form of the RSU was used (RSUE; see Supplementary three RSUs. Dashed black lines represent electrostatic interactions. Thick
Methods). In the absence of ATP (left), the majority of the RSUs (97%) have an dashed green lines represent π–π interactions.
average molecular mass of 63 kDa, corresponding to the RSUE (31.4 kDa) dimer.
Extended Data Fig. 2 | Structure of a catalytic centre in the CSU dimer of the 1 mM ATP and 8.5 μM RSU). These results appear to contradict previous studies
ScAHAS complex. a, Stereo view of the superimposition of the catalytic that show that the KM of pyruvate increases in presence of the ScRSU (2.1 mM for
centres of a CSU dimer in the CSU–RSU complex (light brown) and the CSU alone and 13.3 mM for the complex)7. However, this discrepancy can be
CSU dimer in the absence of the RSU (light blue) (PDB code 5FEM). The explained by the different buffers used in the assay reported here and those
conformation of the active site and the location of ThDP, Mg 2+ and FAD are performed previously7 (Supplementary Methods). Data are mean ± s.e.m. of
unaltered, as well as the binding mode of the herbicide, BSM. FAD, ThDP and technical triplicates. Data were fitted to the Michaelis–Menten equation to
BSM are represented as magenta and orange ball-and-stick models in the obtain the KM values. c, Stereo view of the superimposition of the α–β linker and
structures of the ScAHAS CSU–RSU complex and of the ScAHAS CSU alone, nearby regions in the structures of the ScAHAS complex (light brown) and CSU
respectively. b. KM for pyruvate for the only CSU of ScAHAS (left; assayed using alone (light blue). The dashed line indicates residues that are disordered (not
0.28 μM CSU) and for the ScAHAS complex (right; assayed using 0.14 μM CSU, visible) in the structure of CSU alone.
Article

Extended Data Fig. 3 | See next page for caption.


Extended Data Fig. 3 | Interactions between a ScCSU dimer and Dashed black lines represent electrostatic bonds. The first α-helix (α 1) of the
neighbouring RSUs in the ScAHAS complex. a, Stereo view of a CSU dimer ACT domain in the RSU interacts with the Q-loop of the CSU (chain B) (that is,
and interacting RSUs. The CSU dimer is represented by chains A (dark grey) R101 interacts with Q202) whereas the second α-helix (α 2) interacts with the
and B (light brown), and includes the cofactors FAD (yellow transparent β-domain of the CSU (chain A) (that is, N134 and R137 interact with I413 and
surface) and ThDP (orange transparent surface) and the herbicide BSM P410, respectively). c, Stereo view and magnification of the α–β linker of a CSU.
(blue transparent surface). The RSU dimer is represented by chains C (green) The chains are coloured as in a. Dashed black lines represent electrostatic
and D (magenta). The two bound ATP pairs are shown as a red transparent bonds and white thick dashed lines represent hydrophobic interactions. The
surfaces. Chain G (cyan) and chain P (violet) are RSUs that interact with chain B α–β linker of CSU (chain B) interacts with the ‘contact loop’ of the RSU (chain C),
and chain A, respectively, through their N-terminal extensions. b. Stereo view two β-strands of the ACT domain of the RSU (chain D) and the N-terminal
and magnification of the α-helixes of the ACT domain of RSU (chain C). The extension of the RSU (chain G).
colours of the chains and cofactors are as in a. BSM has been omitted for clarity.
Article

Extended Data Fig. 4 | See next page for caption.


Extended Data Fig. 4 | Regulatory mechanism of ScAHAS. a, The specific comparison of ΔH# (49.2 versus 48.9 kJ/mol) and TΔS # (−16.23 versus −20 .3 kJ/
activity of ScCSU measured at 30 °C as a function of the concentration of mol) values (of CSU + RSU versus CSU, respectively) that account for the value
ScCSU, in the presence (blue data, right y axis) of saturating concentrations of of ΔG# (ΔG# = ΔH# − TΔS #) revealed that the decrease in ΔG# (which is directly
ScRSU (8.5 μM) and ATP (1 mM) or in their absence (black data, left y axis). associated with the increase in kcat) is largely entropy-driven. Data are
Data (technical triplicates) were fitted using Supplementary equation (3) mean ± s.e.m. of the kcat (left) for ΔG# and of the Ea for ΔH# and TΔS #, according to
(Supplementary Methods), yielding a Kd of 0.033 ± 0.005 μM for the CSU in the the literature24. d, e, Effect of the R101A mutation on the activity of ScRSU.
presence of the RSU and ATP, and Kd = 0.63 ± 0.05 μM for the CSU alone. Data are d, Activation of ScAHAS by ScRSUE(WT) or ScRSUE(R101A). The specific activity
mean ± s.e.m. b, Measurement of the energy of activation (Ea) of ScAHAS. of ScAHAS (130 nM in the assay) was measured at different RSU concentrations
Arrhenius plots were generated for the CSU alone (black data, technical with a saturating concentration of ATP (1 mM). The data (technical triplicates)
triplicates) and for CSU in the presence of saturating concentrations of RSU were fitted to the Hill equation, showing that the mutation increases the Kd
and ATP (blue data, technical triplicates) (Supplementary Methods). Data between the CSU and the RSU from 0.27 ± 0.02 (ScRSUE(WT)) to 6 ± 0.3 μM
points between 24 and 36 °C (7 temperatures) were used to perform a linear (ScRSUE(R101A)). Data are mean ± s.e.m. e, The specific activity of the ScCSUs
regression calculation to obtain an Ea value (the absolute value of the slope as a function of their concentration, alone (black data) and in the presence of
multiplied by the gas constant value24 (R = 8.314 kJ/mol)) of 51.7 ± 1.5 kJ/mol for 1 mM ATP and 2.6 μM ScRSUE(WT) (blue data, saturating concentration) or
the CSU in presence of the RSU and ATP, and Ea = 51.5 ± 2.1 kJ/mol for the CSU 20 μM ScRSUE(R101A) (red data, saturating concentration). The data (technical
alone. Data are mean ± s.e.m. (s.e.m. of the slope value multiplied by R). triplicates) were fitted to Supplementary equation (3) (Supplementary
c, Measurement of the thermodynamic parameters of the activation at 30 °C. Methods), yielding a Kd of 0.04 ± 0.005 and 0.036 ± 0.007 μM for the CSU in the
Left, the kcat of ScCSU at 30 °C was measured in the presence (blue data, right presence of RSUE(WT) and RSUE(R101A), respectively. For the CSU alone, the Kd
y axis) of saturating amounts of ScRSU (33 μM) and ATP (1 mM) or in their is 0.62 ± 0.04 μM confirming the value shown in a. The values of kcat—which are
absence (black data, left y axis). Data points are technical replicates (n = 5). Data also derived from Supplementary equation (3)—are 11.1 ± 0.2, 40.9 ± 1.1 and
are mean ± s.e.m. of the five replicates. Right, values for the thermodynamic 19.5 ± 0.8 s−1 for the CSU alone, the CSU with RSUE(WT) and the CSU with
parameters of the activation at 30 °C. The values of the change in Gibbs free RSUE(R101A), respectively, showing that the R101A mutation strongly affects
energy (ΔG#), enthalpy (ΔH#), and temperature and entropy (TΔS #) were catalysis. Data are mean ± s.e.m.
calculated using the equations derived from the transition state theory24. The
Article

Extended Data Fig. 5 | See next page for caption.


Extended Data Fig. 5 | Model for the regulation of the ScAHAS reaction. through their N- and C-terminal extensions, which transmit the movements
A hybrid image showing the catalytic centres taken from an asymmetric ScCSU between catalytic centres (red arrows). When the complex is formed, the RSU
dimer in the presence of pyruvate (grey and brown) (PDB code 5INU) and works at two levels. (1) The RSU creates a ‘rigid scaffold’ (magenta) that is
the α 1-helices (pink and green) of the RSU from the complex. The different formed by interactions of the ACT regions (that is, the α 2-helix and two
conformations of FAD and the Q-loops, and the He-ThDP intermediate or ThDP β-strands of the ACT domain) with the α- and β-domains and the α–β linker of
are shown. The Q-loops and their N-terminal and C-terminal extensions have the CSUs (Extended Data Fig. 3). This restrict movements of the CSU that is not
been drawn in brown and dark grey for catalytic centre 1 and 2, respectively. involved in the communication between active sites. (ii) The α 1-helices of the
The model shows how the conformational change of the active sites during the RSUs create a bridge (α 1-bridge) between Q-loops (curve line in magenta) that
catalytic cycle20 triggers a movement of the Q-loops that is complementary increases the efficiency of the mechanical transmission between catalytic
between the catalytic centres (red arrows). This feature facilitates the centres. Both aspects are involved in decreasing the entropy of the enzyme–
synchronization of the reactions that occur in the two catalytic centres of a substrate complex before activation, leading to an increase in kcat.
CSU dimer13. In the ScCSU dimer, direct contact between the Q-loops is made
Article

Extended Data Fig. 6 | See next page for caption.


Extended Data Fig. 6 | Binding site interactions, cryo-EM map for l-valine in light blue for structures of the complex in the absence and presence of l-valine,
AtRSU and consequences of l-valine binding. a, Cryo-EM densities of the two respectively. The RSUs are shown in dark green and magenta for the repeats in
BCAA-binding sites (site 1 and 2) occupied by l-valine. l-Valine is shown as a the structure of the complex in the absence of l-valine, and cyan and coral for
green stick model. The cryo-EM map (5.5σ) is overlaid. b, Stereo view of the the repeats in the structure of the complex with l-valine bound. The cofactors
interaction of l-valine in site 1 with the α 1-helices of the ACT domains. Dashed FAD (yellow) and ThDP (orange) are shown in the structure of the l-valine free
lines represent hydrogen bonds (black), van der Waals interactions (blue) and complex only. l-Valine (ball and stick representation) has a light-yellow
hydrophobic interactions (yellow). c, Superimposition of the free AtAHAS transparent surface around the outside. The comparison of the structures
complex and the l-valine-bound complex, representing one of the four CSU shows that the CSU dimer expands when l-valine is bound to the ACT domains
RSU pairs that interact with each other. The CSU dimers are shown in green and of the RSUs.
Article

Extended Data Fig. 7 | Structural changes in the ACT domains of AtRSU that R110–D204 salt bridge. b, Stereo view showing the conformational change in α 2
are triggered by the binding of l-valine. a, Stereo view of the translocation of caused by the bending of α 1. The conformational change of α 1 induced by
R110 induced by the binding of l-valine in site 1 of the ACT domain. When l-valine triggers new interactions to occur between α 1 and α 2 (that is, the
l-valine is present the backbone of the α 1-helix of repeat 1 is shifted (black interaction of V138 (α 2) with R110 (α 1) and L139 (α 2) with Y112 (α 1)) leads to the
arrows), causing a marked translocation of R110 and the breaking of the bending of α 2). The colour scheme is as described in Extended Data Fig. 6c.
Extended Data Fig. 8 | See next page for caption.
Article
Extended Data Fig. 8 | The structure of the fungal hybrid complex formed RSU alone). The amide group of l-valine forms either two or three hydrogen
by the CnCSU and ScRSU in the presence of l-valine and comparisons with bonds to the RSU (N113 side chain, I114 carbonyl oxygen, D328 side chain in
the AtAHAS complexes. a, The overall structure of the hybrid complex AtAHAS; N104 side chain, I105 carbonyl oxygen in Cn–ScAHAS; N37 side chain,
resembles the ScAHAS and AtAHAS complexes, possessing the characteristic V38 carbonyl oxygen, Q19 side chain in E. coli RSU alone). The side chain
Maltese cross shape. b, The addition of ATP leads to the activation of the fungal of l-valine is stabilized by hydrophobic and/or van der Waals interactions with
hybrid complex, whereas the addition of l-valine inhibits its activity. This nearby hydrophobic side chains. The similarity of the network of interactions
experiment was performed once. c, Left, the cryo-EM map at 3.4 Å resolution involving the binding of l-valine that is observed in all three complexes,
(8σ level) in the AtAHAS complex with l-valine overlaid and optimally fitted. and the fact that these three sites align when the sequences are compared
Middle, the hybrid Cn–ScAHAS complex with the Fo − Fc electron density (Supplementary Fig. 2) shows that the binding mode of l-valine is conserved in
(3.5σ level) with l-valine overlaid and optimally fitted. Right, the binding mode bacterial, fungal and plant AHASs. d, Comparison of the conformational
of l-valine to E. coli RSU (structure determined in the absence of the CSU at 2.3 Å changes that occur in the AtAHAS and in the Cn–ScAHAS complexes on l-valine
resolution; PDB code 5YPW). In all three complexes, the carboxylate group binding. These changes are highly conserved in the two complexes as
of l-valine forms three hydrogen bonds to backbone amides of the RSUs (I114, emphasized by the arrows.
V332, L333 in AtAHAS; I105, V90, L91 in Cn–ScAHAS; V38, V23, M24 in the E. coli
Extended Data Table 1 | Data collection and refinement statistics for ScAHAS complexes

Data collection and refinement statistics, including molecular replacement analyses, for the ScAHAS CSU–RSU and CnCSU–ScRSU (AHAS) + valine complexes. One crystal was used to obtain
each dataset.
*Values in parentheses are for the highest-resolution shell.
Article
Extended Data Table 2 | Cryo-EM data collection, refinement and validation statistics for the AtAHAS complexes
nature research | reporting summary
Corresponding author(s): Luke William Guddat/ Thierry Lonhienne
Last updated by author(s): 2nd October 2019

Reporting Summary
Nature Research wishes to improve the reproducibility of the work that we publish. This form provides structure for consistency and transparency
in reporting. For further information on Nature Research policies, see Authors & Referees and the Editorial Policy Checklist.

Statistics
For all statistical analyses, confirm that the following items are present in the figure legend, table legend, main text, or Methods section.
n/a Confirmed
The exact sample size (n) for each experimental group/condition, given as a discrete number and unit of measurement
A statement on whether measurements were taken from distinct samples or whether the same sample was measured repeatedly
The statistical test(s) used AND whether they are one- or two-sided
Only common tests should be described solely by name; describe more complex techniques in the Methods section.

A description of all covariates tested


A description of any assumptions or corrections, such as tests of normality and adjustment for multiple comparisons
A full description of the statistical parameters including central tendency (e.g. means) or other basic estimates (e.g. regression coefficient)
AND variation (e.g. standard deviation) or associated estimates of uncertainty (e.g. confidence intervals)

For null hypothesis testing, the test statistic (e.g. F, t, r) with confidence intervals, effect sizes, degrees of freedom and P value noted
Give P values as exact values whenever suitable.

For Bayesian analysis, information on the choice of priors and Markov chain Monte Carlo settings
For hierarchical and complex designs, identification of the appropriate level for tests and full reporting of outcomes
Estimates of effect sizes (e.g. Cohen's d, Pearson's r), indicating how they were calculated
Our web collection on statistics for biologists contains articles on many of the points above.

Software and code


Policy information about availability of computer code
Data collection All of the open source code used has been cited in the paper. No original software was developed.

Data analysis All of the software used in data analysis performed is cited in the paper. No original software was developed.
For manuscripts utilizing custom algorithms or software that are central to the research but not yet described in published literature, software must be made available to editors/reviewers.
We strongly encourage code deposition in a community repository (e.g. GitHub). See the Nature Research guidelines for submitting code & software for further information.

Data
Policy information about availability of data
All manuscripts must include a data availability statement. This statement should provide the following information, where applicable:
- Accession codes, unique identifiers, or web links for publicly available datasets
- A list of figures that have associated raw data
- A description of any restrictions on data availability

The coordinates and structure factors for the ScAHAS complex and the coordinates and cryo-EM map for the AtAHAS complex have been deposited in the
Protein Data Bank (www.pdb.org) with the PDB ID codes of 6U9D and 6U9H, respectively. The cryo-EM maps have also been submitted to the Electron
October 2018

Microscopy Data Bank (EMDB) under the accession code EMD-20700. Raw data figure: 2b ext dat 1b, 2b, 5, 9b, supp fig 3, 5, 9, 13, 14

1
nature research | reporting summary
Field-specific reporting
Please select the one below that is the best fit for your research. If you are not sure, read the appropriate sections before making your selection.
Life sciences Behavioural & social sciences Ecological, evolutionary & environmental sciences
For a reference copy of the document with all sections, see nature.com/documents/nr-reporting-summary-flat.pdf

Life sciences study design


All studies must disclose on these points even when the disclosure is negative.
Sample size not applicable

Data exclusions not applicable

Replication not applicable

Randomization Not applicable

Blinding Not applicable

Reporting for specific materials, systems and methods


We require information from authors about some types of materials, experimental systems and methods used in many studies. Here, indicate whether each material,
system or method listed is relevant to your study. If you are not sure if a list item applies to your research, read the appropriate section before selecting a response.

Materials & experimental systems Methods


n/a Involved in the study n/a Involved in the study
Antibodies ChIP-seq
Eukaryotic cell lines Flow cytometry
Palaeontology MRI-based neuroimaging
Animals and other organisms
Human research participants
Clinical data

October 2018

You might also like