Main

We purified and crystallized a truncated short Mgm1 isoform from the thermophilic fungus C. thermophilum (hereafter denoted Mgm1) (Fig. 1a, Extended Data Fig. 1a, Supplementary Fig. 1). Crystals of this construct grown in the absence of nucleotides diffracted to a resolution of 3.6 Å. The structure was solved by single anomalous dispersion (Extended Data Fig. 1b, c, Extended Data Table 1).

Fig. 1: The structure of Mgm1 reveals a paddle domain that is required for membrane binding.
figure 1

a, Domain and isoform architecture of Mgm1. B, bundle signalling element; MSS, mitochondrial signal sequence; TM, transmembrane domain. l and s denote long and short isoforms of Mgm1, respectively. b, Ribbon representation of Mgm1. Domains are coloured individually as in a. The inset shows the disulfide bond between the conserved cysteine residues C812 and C821 and the conserved positively charged residues R748 and K749 in the paddle domain. Note that C821 is in the centre of the paddle whereas C812 is closer to its periphery. Apart from the loss of the disulfide bridge, the C821S mutation may therefore disrupt the paddle conformation more strongly than the C812S mutation. c, Representative liposome-binding experiment. P, pellet; S, supernatant; WT, wild type. n = 4 independent experiments. d, e, GTPase activity (d) (n = 4 independent experiments; data are mean ± s.e.m.) and negative-stain electron micrographs (e) of liposome tubulation of wild type Mgm1 and indicated mutants (n = 2 independent experiments; scale bars, 50 nm). Quantification of all experiments is shown in Extended Data Fig. 3a and the raw data is available in Supplementary Fig. 2.

Source data

Mgm1 contains four domains: a G domain, a bundle signalling element (BSE) domain, a stalk and a paddle domain (Fig. 1a, b). The G domain closely resembles that of human dynamin (Extended Data Fig. 2). An interface across the nucleotide-binding site (the ‘G interface’), which is responsible for dimerization of G domains in the dynamin superfamily, is highly conserved in Mgm1 (Extended Data Fig. 1e). The adjacent BSE domain consists of three helices that are derived from different regions of Mgm1 (Fig. 1a, b). The BSE domain forms contacts with the G domain, as is the case in the closed conformation of dynamin17,18,19. The Mgm1 stalk domain is shorter than the dynamin stalk, comprising an antiparallel kinked four-helix bundle (Fig. 1b, Extended Data Fig. 2c, d). Unique to Mgm1 is the paddle domain, an elongated three-helix domain at the tip of the stalk, which is inserted between stalk helices α3S and α4S and contains a disulfide bridge that links Cys812 to Cys821 (Fig. 1b). Dynamin has a membrane-binding pleckstrin homology domain in the corresponding position.

We mutated two positively charged, surface-exposed residues as well as the two cysteines of the disulfide bridge in the paddle domain. In co-sedimentation experiments, the Mgm1 construct efficiently bound to Folch liposomes (lipids from bovine brain), whereas the mutant proteins bound less strongly (Fig. 1c, Extended Data Fig. 3a). Mgm1 bound to the non-hydrolysable GTP analogue GTPγS with a dissociation constant (Kd) of 9 µM (Extended Data Fig. 3b). The intrinsic GTPase activity of Mgm1 (about 0.5 min−1 at 37 °C) was accelerated about 500-fold in the presence of Folch liposomes, reaching rates of 270 min−1. Yeast Mgm1 and human OPA1 also show increased GTPase activity in the presence of liposomes20,21. Stimulation of GTPase activity was considerably lower for all paddle mutants (Fig. 1d).

When incubated with liposomes, Mgm1 induced tubulation and coated the membrane surface in a regular pattern (Fig. 1e), with or without added nucleotide (Extended Data Fig. 3e, f). The membrane-remodelling activity of the paddle mutants was reduced, indicating that the paddle constitutes a membrane-binding site (Fig. 1e, Extended Data Fig. 3g).

The asymmetric unit of the crystals contained an Mgm1 dimer. The dimer interface (termed interface-2, in analogy to that of dynamin17,18) includes a hydrophobic core that is flanked by polar residues (Fig. 2a). Using analytical ultracentrifugation, we detected a concentration-dependent monomer–dimer equilibrium for Mgm1 (Fig. 2b, Extended Data Fig. 3c). The F840D mutation, in the centre of the hydrophobic dimer interface, rendered the protein monomeric. Assembly via the Mgm1 stalk interface-2 results in a V-shaped dimer, whereas dynamin stalks form an X-shaped dimer (Extended Data Fig. 2d, e).

Fig. 2: The Mgm1 dimer.
figure 2

a, The dimer is formed via interface-2 between opposing stalks. b, Sedimentation velocity analysis of Mgm1 and Mgm1(F840D) at different protein concentrations. The unusually small apparent sedimentation coefficients (sapp, expressed in Svedberg) of both the monomeric and the dimeric species are consistent with their non-globular structures. The data for Mgm1 can be fitted to a Kd of 1 µM. n = 1. c, Negative-stain electron micrographs of tubulated liposomes with wild-type Mgm1 and with Mgm1(F840D), in which an interface-2 residue is mutated. n = 2 independent experiments; scale bars, 50 nm. d, e, Liposome binding (d) and GTPase activity (e) for proteins with the stated interface-2 mutations (n = 4 independent experiments). For e, data are mean ± s.e.m.

Source data

Mutation of several interface-2 residues reduced both the binding of Mgm1 to liposomes and the extent to which its GTPase activity was stimulated in the presence of liposomes (Fig. 2d, e, Extended Data Fig. 3a). The F840D mutation had the most severe effect: Mgm1(F840D) failed to tubulate liposomes and did not form a regular protein pattern (Fig. 2c–e, Extended Data Fig. 3g). These results confirm the importance of interface-2 for the assembly of Mgm1 on the membrane surface.

We used a yeast model system to express Mgm1 mutants in a strain in which the expression of endogenous Mgm1 was under the control of the galactose-inducible and glucose-repressed GAL1 promoter (Extended Data Fig. 4a). Loss of Mgm1 expression was associated with a rapid and irreversible loss of the mitochondrial genome, fragmentation of the mitochondrial network and the subsequent inability to switch to respiratory metabolism upon glucose depletion6,10 (Extended Data Fig. 4b, c). Re-expression of yeast Mgm1 rescued the loss of mitochondrial respiratory function, as assessed by yeast growth, the presence of mitochondrially encoded cytochrome c oxidase 1 protein (Cox1) and restoration of the mitochondrial network (Extended Data Fig. 4b–g). Consistent with the liposome-binding assays, yeast Mgm1(F805D) (corresponding to F840D in C. thermophilum Mgm1)—but not Mgm1(N675A) (corresponding to I700D in C. thermophilum Mgm1, Supplementary Fig. 1)—failed to complement the loss of endogenous Mgm1 (Extended Data Fig. 4d, e, g). These results highlight the importance of interface-2 for Mgm1-dependent maintenance of mitochondrial DNA and respiration-competent mitochondria.

In the crystals, two Mgm1 dimers assembled into a tetramer via another stalk interface of approximately 1,000 Å2; in further analogy to dynamin, we refer to this as stalk interface-1 (Fig. 3a). The tetramer is additionally stabilized by a contact of approximately 1,100 Å2 between the BSE domain of one dimer and the stalk of the adjacent dimer. Interface-1 and the BSE–stalk contact site are highly conserved in the Mgm1 family (Extended Data Fig. 1e). Notably, the interface-1 interaction induces a bend of 20° between two stalk dimers (Extended Data Fig. 2e).

Fig. 3: Assembly mechanism of Mgm1.
figure 3

a, The tetramer, as seen in the crystal. Two dimers (grey or colour-coded by domain) interact via stalk interface-1 and a small BSE–stalk contact. See insets for details. b, Yeast respiratory growth complementation assays with Mgm1 containing mutations in interface-1 residues (left; D542 and K545 in yeast Mgm1 correspond to D559 and K562 in the C. thermophilum protein) and residues of the BSE–stalk contact (right; Y520 and R621 in yeast Mgm1 correspond to Y537 and R646 in the C. thermophilum protein). A representative growth curve from n = 3 independent experiments is shown. EV, empty vector control; OD600nm, optical density at 600 nm; AU, arbitrary units. See also Extended Data Fig. 4.

Mgm1(D559A) or Mgm1(K562A), containing mutations in interface-1, and Mgm1(Y537A) or Mgm1(R646A), containing mutations in the BSE–stalk contact, did not show major differences in liposome binding or GTPase activity compared to Mgm1 (Extended Data Fig. 3a, d). These mutant proteins also tubulated liposomes and formed a regular pattern on the membrane (Extended Data Fig. 3g). However, when these mutations were introduced into the corresponding positions of yeast Mgm1, the resultant mutant proteins failed to complement the loss of wild-type Mgm1 in respiratory growth (Fig. 3b), mitochondrial genome maintenance and mitochondrial morphology (Extended Data Fig. 4d, g). Notably, yeast Mgm1(Y520A)—containing a mutation in the BSE–stalk contact in the tetramer interface—exerted a strong dominant negative effect on respiratory yeast growth when co-expressed with endogenous Mgm1 (Extended Data Fig. 4h). The corresponding yeast strain retained mitochondrial DNA (Extended Data Fig. 4i), enabling us to examine Mgm1-specific deficits on mitochondrial morphology and ultrastructure. The expression of yeast Mgm1(Y520A) induced fragmentation of the mitochondrial network (Extended Data Fig. 4j, k), a reduction in the number and length of cristae and an increase in crista diameter (Extended Data Fig. 4l, m).

We used electron cryo-tomography (cryo-ET) and subtomogram averaging to determine the structure of membrane-bound Mgm1 (Fig. 4a, b, Extended Data Fig. 5a, b). In the absence of nucleotide, or upon the addition of GTPγS, Mgm1 remodelled Folch liposomes into membrane tubes of varying diameters ranging from around 18 nm to 140 nm. The Mgm1 coat decorated membrane tubes in a regular lattice. For subtomogram averaging, preformed tubes with diameters of around 20 nm were used in order to increase the number of particles for averaging. These tubes also stimulated the GTPase activity of Mgm1, although less strongly than did Folch liposomes (Extended Data Fig. 5c). The final resolution of the subtomogram average volume was 14.7 Å for the nucleotide-free and the nucleotide-bound forms (Extended Data Table 1). No substantial differences were apparent between the two volumes. Notably, the Mgm1 tetramer fits the subtomogram average volume with only minor positional domain rearrangements (Fig. 4b, Extended Data Fig. 6a, c, e). The G domain was in a closed conformation relative to the BSE domain and was located furthest from the membrane, the stalk was in the middle, and the paddle was next to the membrane. The Mgm1 coat in Fig. 4b, c can be viewed as a left-handed four-start helix, consisting of four parallel helical filaments (Extended Data Fig. 7a, b). Similar filaments were observed on Folch lipid tubes of different diameters, although their helical parameters varied (Extended Data Fig. 5d). The filament backbone was formed by stalks oligomerizing in an alternating fashion via interface-1 and interface-2. This is in contrast to dynamin filaments, in which the stalks oligomerize via three interfaces22,23 (Extended Data Fig. 2e). Another difference compared with dynamin22 is that we did not observe interactions between the G domains of adjacent helix turns. Instead, contact was established through the paddle domains (Fig. 4b, c, Extended Data Fig. 6a). Mutation of the conserved residues F779 and S780 in the paddle contact site affected membrane binding and stimulated GTPase activity only mildly (Extended Data Fig. 3a, d). Expression of the corresponding mutant protein in yeast complemented the loss of endogenous Mgm1 with respect to respiratory growth, but the cells exhibited moderate alterations of mitochondrial morphology and mitochondrial genome maintenance (Extended Data Fig. 4d, f, g).

Fig. 4: Mgm1 forms a helical lattice on the outside of lipid tubes.
figure 4

a, Mgm1 forms a regular protein coat on the outer surface of galactocerebroside-containing lipid tubes, enabling analysis by cryo-ET. b, The subtomogram average shows the protein lattice with Mgm1 flexibly fitted into the cryo-ET volume of the apo form. The outer leaflet of the membrane is not well defined in this reconstruction. Arrows indicate density not attributable to the protein, which was assigned to the outer membrane leaflet. c, Four filaments of Mgm1 dimers wrap around a membrane tube in a left-handed surface lattice. Stalks assemble via interface-1 and interface-2 in an alternating manner (Extended Data Fig. 7). d, Tube-pulling assay for the generation of a tube surface that is accessible from the outside. Mgm1 (green fluorescence) binds to the tube and the GUV surface (n = 8 independent experiments).

The tendency for Mgm1 to form a left-handed helix on the convex exterior of membrane tubes is consistent with the curvature of the crystallographic tetramer that arises from the interaction between interfaces-1 of two dimers. A model in which several dimers are connected via identical interfaces-1 results in a continuous filament with dimensions and helix parameters (radius, pitch) similar to those observed by cryo-ET (Extended Data Fig. 7b, c). Microsecond-scale molecular dynamics simulations starting from the crystallographic tetramer provide further evidence for the curvature preference of the Mgm1 interface-1 (Extended Data Fig. 7d–k). The most likely curvature and twist in the simulated interface-1 was the same as that in the crystal lattice. The simulation results also suggested that there is sufficient flexibility in interface-1 to account for the observed variable radii of Mgm1 helical filaments.

We followed the dynamics of Mgm1 assembly on the membrane by live fluorescence confocal imaging. By manipulating streptavidin beads adhering to giant unilamellar vesicles (GUVs) with optical tweezers24, membrane tubes can be pulled out of the GUV in a controlled manner. Mgm1 was then injected into the chamber with a second pipette (Fig. 4d).

Consistent with the results of cryo-ET and negative-stain electron microscopy, Mgm1 adapted to different degrees of membrane curvature by decorating the outer surface of the membrane tube and the GUV. The addition of GTP after assembly did not result in membrane scission under these conditions, but the force required to hold the tube in place (measured as a function of bead displacement) increased by a factor of three to five. This is consistent with a GTP-dependent structural rearrangement of the Mgm1 coat and/or a GTP-dependent expansion of the membrane tube (Extended Data Fig. 8a, b).

Cryo-ET analysis revealed that Mgm1 occasionally decorated the inside surface of Folch membrane tubes in a regular pattern, which suggests that the liposomes were leaky (Fig. 5a, Extended Data Fig. 5e). In further experiments, liposomes were sonicated for a few seconds after the addition of Mgm1 to promote formation of the internal lattice. Tubes with an internal lattice were much wider and were less variable in diameter (range 90–105 nm) (Extended Data Fig. 5e). The negative (concave) membrane curvature on the inner surface of a larger tube resembles the inside of mitochondrial cristae. Subtomogram averages of Mgm1 decorating the inner vesicle surface were obtained for the nucleotide-free and the GTPγS-bound form (Fig. 5, Extended Data Fig. 5f, g, Extended Data Table 1). At an estimated resolution of 20.6 Å for the nucleotide-free and 18.8 Å for the nucleotide-bound form, the subtomogram average volumes appeared very similar.

Fig. 5: Mgm1 forms a lattice on the inside of lipid tubes.
figure 5

a, Mgm1 in the apo form decorates the inner surface of lipid tubes. The subtomogram average reveals regular protein arrays on the inner surface of membrane tubes; a magnified view shows Mgm1 flexibly fitted into the cryo-ET volume, with the G domains dimerized via the G interface. b, The Mgm1 lattice on the inner membrane surface of a tube. Interface-1 between tetramers is open, see also Extended Data Fig. 6. c, Tube-pulling assay for the generation of a membrane tube that is accessible from the inside. Red indicates lipid fluorescence. Mgm1 (green fluorescence) binds to the neck and the inner surface of the tube. The arrow points to the entry of the membrane tube (n = 7 independent experiments).

As in the external lattice, the crystallographic Mgm1 tetramer fitted the subtomogram average volume of the internal lattice with only minor rearrangements (Fig. 5a, b, Extended Data Fig. 6b, d, e). The G domains were furthest from the membrane facing into the tube, the stalk was in the middle and the paddle domain was next to the membrane. The arrangement of tetramers on the internal lattice differed markedly from that on the external membrane surface (Fig. 5a, b, Extended Data Fig. 6a–e). Rather than through interface-1, assembly involved a contact between neighbouring tetramers that included conserved patches in the BSE and stalk domains, closely resembling the linear arrangement of tetramers in the crystal lattice (Extended Data Fig. 6f). The angle between filaments of Mgm1 tetramers and the plane perpendicular to the tube axis was 69°, whereas it was 21° in the external lattice (Fig. 5b, Extended Data Fig. 6a, b). Another major difference compared with the external lattice is that the G domains were in close contact, and their orientation indicated that interaction occurred through the G interface. This G domain contact was enabled by the opening of interface-1, even though the G domain/BSE interface remained closed. As with Mgm1 filaments on the surface of the outer membrane, the paddle domains contributed to lattice formation.

To investigate the assembly of Mgm1 on negatively curved membrane surfaces, streptavidin beads were pulled inside a GUV (Fig. 5c). In this situation, Mgm1 assembled preferentially at the funnel-shaped connection between the tube and the GUV and then grew further into the tube. Mgm1 did not redistribute on the membrane in the presence of GTP (Extended Data Fig. 8c). However, as with the positively curved (convex) membranes, the force on the tube increased in a GTP-dependent manner (Extended Data Fig. 8d). Together with the results from cryo-ET, the results of these experiments demonstrate conclusively that Mgm1 can form stable assemblies on negatively curved membranes.

Our study reveals the structural basis of Mgm1 assembly—via the stalks—into dimers, tetramers and helical filaments. Dynamin17,18,22, dynamin-like MxA25 and DNM1L26,27 are likewise known to oligomerize via their stalks into helical filaments, although important parameters of the assembly mode differ (Extended Data Fig. 2e). In dynamin, the G domains of adjacent turns transiently dimerize and mediate a GTPase-dependent power stroke19,28, which is thought to pull the filament turns against each other29,30. We propose that Mgm1 may undergo a similar power stroke, for the following reasons: the G domains and BSE domains of Mgm1 and dynamins are virtually identical; the mechanisms of membrane-stimulated GTPase activity are similar; the G domains in our cryo-ET reconstructions of Mgm1 are closely apposed; a GTP-dependent force was observed in the tube-pulling assays; and temperature-sensitive mutations in Mgm1 localize to one of the GTP-binding loops (switch I), the G/BSE domain interface or the assembly interface-110 (Extended Data Fig. 1d). Furthermore, the GTPase activity of OPA1 is required to sustain cristae morphology14.

A dynamin-like power stroke would result in different remodelling processes depending on the assembly geometry of the Mgm1 filaments. When assembled on positively curved membranes in a left-handed helical pattern (Fig. 4), a dynamin-like power stroke would expand the diameter of the lipid tube. Conversely, a right-handed helix pattern would result in constriction, as observed in dynamin (Extended Data Fig. 9a, b, Supplementary Video 1).

Mgm1 is the only known member of the dynamin superfamily that can assemble on the inside of membrane tubes—a membrane geometry similar to that of mitochondrial inner membrane cristae. We postulate that Mgm1 can form helical filaments at the inside of membrane tubes with the shape and dimensions of crista junctions (Extended Data Fig. 9c, d). Notably, a power stroke in a left-handed helical assembly on the inside of a membrane tube would constrict its diameter—as observed for the crista junctions upon OPA1 overexpression14—whereas a right-handed assembly would expand it (Extended Data Fig. 9a, b, Supplementary Video 1). In Extended Data Fig. 9e–h, we suggest how the membrane geometry of different filament assemblies might explain inner-membrane fusion, scission, or the stabilization of cristae.

Taken together, our structural and functional studies reveal the molecular basis of the assembly of Mgm1 into filaments, and provide models of how the rearrangements of these filaments induce remodelling of the inner mitochondrial membrane.

Methods

Data reporting

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

Protein expression and purification

C. thermophilum Mgm1 (Mgm1, amino acids 219–912) and indicated mutants of this construct were expressed from pET46-EK/LIC (Novagen) as N-terminal His6-tag fusion followed by a PreScission cleavage site. Proteins were expressed in Escherichia coli host strain BL21-DE3, and bacteria were cultured in TB medium at 37 °C followed by induction with 200 µM isopropyl-β-d-thiogalactopyranoside and a temperature shift to 20 °C for overnight expression. Selenomethionine-substituted Mgm1 was expressed in M9 minimal medium, supplemented with l-amino acids lysine, phenylalanine, threonine (100 mg l−1), isoleucine, leucine, valine and selenomethionine (50 mg l−1), using the same vector and host strain as for native protein expression31. Cells were resuspended in buffer A (25 mM HEPES/NaOH (pH 7.8), 350 mM NaCl, 150 mM KCl, 2 mM MgCl2, 1 µM DNase (Roche), 500 µM Pefabloc (Roth)) and disrupted by a microfluidizer (Microfluidics). Cleared lysates (95,000g, 1 h, 4 °C) were incubated with Benzonase (Novagen) for at least 30 min at 4 °C before application to a Co2+-Talon column (Clontech). Proteins were eluted with buffer A containing an additional 100 mM imidazole. Fractions containing Mgm1 were incubated with 2.4 mM β-mercaptoethanol and His6-tagged Prescission protease overnight at 4 °C. Then, imidazole, β-mercaptoethanol and the free His-tag were removed by using 50 kDa molecular weight cut-off concentrators (Amicon) and washing with buffer A, before a second application to a Co2+-Talon column to remove non-cleaved His-tagged Mgm1 and protease. The flow-through and four column volumes of washing buffer A were collected and concentrated. Finally, Mgm1 was purified by size-exclusion chromatography on a Superdex200 column (GE) in buffer A. Fractions containing Mgm1 were pooled, concentrated and flash-frozen in liquid nitrogen (Extended Data Fig. 1a). Selenomethionine-substituted protein and mutant proteins were purified using the same protocol.

Crystallization and structure determination

Crystallization trials using the sitting-drop vapour-diffusion method were performed at 4 °C using a Gryphon pipetting robot (Art Robbins Instruments) and Rock Imager storage system (Formulatrix). 300 nl of the selenomethionine-substituted Mgm1 at a concentration of 12.9 mg ml−1 was mixed with an equal volume of reservoir solution containing 8% PEG400, 3% isopropanol, 100 mM Na-citrate buffer (pH 5.5). Crystals appeared after 2 weeks and had final dimensions of 500 µm × 200 µm × 50 µm. During flash-cooling of the crystals in liquid nitrogen, a cryo-solution containing additionally 20% ethylene glycol was used. The dataset was recorded at BL14.1 at BESSY II, Berlin. One native dataset was collected at a wavelength of 0.9794 Å and a temperature of 100 K from a single crystal and processed and scaled using the XDS program suite32,33. Twenty-two out of twenty-six Se sites were detected by Autosol/PHENIX34 for two molecules in the asymmetric unit (80% solvent content). The density showed a continuous trace for the peptide backbone and clear anomalous signals for the positions of the selenomethionine side chains. The initial model was built by adapting the BSE and stalk domain from the human dynamin 3 structure (Protein Data Bank (PDB): 5A3F) to the density. For chain A, the nucleotide-free G domain of human dynamin 1 (PDB: 2AKA) fitted the density well, whereas density for the G domain of chain B was weak. The G domain was therefore omitted in the initial chain B model. The electron density for loop 2 in the stalk (L2S, see Supplementary Fig. 1) and for the paddle was well defined. Building of missing residues in this area was guided by the anomalous signal of the selenomethionine side chains. The model was built using COOT35 and iteratively refined with Phenix 1.11.1-257536 including Hendrickson–Lattman coefficients, non-crystallographic symmetries of the respective domains, secondary structure restraints, one TLS group per domain and one B factor per amino acid. Occupancy of side chains with considerable radiation damage was reduced to 0.8 or 0.6 for surface exposed glutamate or aspartate residues, and to 0.8 or 0.5 for selenomethionine residues. Finally, the G domain from chain A was transplanted to chain B and refined as a rigid body. Two ethylene glycol molecules were built into remaining difference density at the end of the refinement. 1,252 residues out of 1,304 refined residues (96%) are in the most favoured regions of the Ramachandran plot and 3 residues are in the disallowed regions (0.23%), as analysed with Phenix. Buried surface areas were calculated using the PISA server37. Domain superpositions were performed with lsqkab from the CCP4 program suite38. Figures were prepared with the PyMOL Molecular Graphics System, Version 2.0 (Schrödinger, LLC.). Sequences were aligned using CLUSTAL Omega39 and adjusted by hand.

Analytical ultracentrifugation experiments

All measurements were performed in 25 mM HEPES/NaOH pH 7.8, 50 mM NaCl, 150 mM KCl and 1 mM MgCl2 at 20 °C using an Optima XL-A centrifuge (Beckman) and an An50Ti rotor equipped with double sector cells. Depending on protein concentration, the distribution of the protein in the cell was monitored at 230 or 280 nm. Data were analysed using the software SedFit40. Sedimentation velocity was run at 40,000 r.p.m. for 3 h, sedimentation equilibrium was performed at 8,000 r.p.m.

Liposome co-sedimentation assays

Liposomes were prepared as previously described (www. endocytosis.org). Folch liposomes (0.6 mg ml−1) (total bovine brain lipids fraction I from Sigma) in 25 mM HEPES/NaOH (pH 7.8), 60 mM NaCl, 100 mM KCl and 0.5 mM MgCl2 were incubated at room temperature with 4 µM of the indicated Mgm1 construct for 10 min in 40 µl reaction volume, followed by spinning at 210,000g for 10 min at 20 °C and SDS–PAGE analysis of the supernatant and the pellet. For quantification, the protein bands were integrated using ImageJ and the intensity of each band (supernatant or pellet) was divided by the sum of the intensities from supernatant and pellet.

Isothermal titration calorimetry

Isothermal titration calorimetry experiments were performed at 18 °C in a PEAQ-ITC (Malvern) in 20 mM HEPES/NaOH pH 7.5, 60 mM NaCl, 100 mM KCl and 0.5 mM MgCl2, with 50 µM Mgm1 in the reaction chamber and 1 mM GTPγS in the syringe. Malvern software was used to integrate the binding isotherms and calculate the binding parameters.

GTP hydrolysis assay

GTPase activities of 1 µM of the indicated Mgm1 constructs were determined at 37 °C in 25 mM HEPES/NaOH (pH 7.8), 60 mM NaCl, 100 mM KCl and 0.5 mM MgCl2, in the absence and presence of 0.1 mg ml−1 Folch liposomes, using saturating concentrations of GTP as substrate (1 mM for the basal and 3 mM for the stimulated reactions). Reactions were initiated by the addition of protein to the reaction. At different time points, reaction aliquots were diluted 15-fold with GTPase buffer and quickly frozen in liquid nitrogen. Samples were analysed with an HPLC system (Agilent Technologies). Denatured proteins were adsorbed to a C18 guard column and nucleotides were separated via a reversed-phase Hypersil ODS-2 C18 column (250 × 4 mm), with 10 mM tetrabutylammonium bromide, 100 mM potassium phosphate (pH 6.5), 7.5% acetonitrile as running buffer. Nucleotides were detected by absorption at 254 nm and quantified by integration of the corresponding peaks. Rates were derived from a linear fit to the initial reaction.

Negative-stain electron microscopy

For electron microscopy of negatively stained samples in a Zeiss EM910, 4 µM Mgm1 (amino acids 219–912) in 25 mM HEPES/NaOH (pH 7.8), 60 mM NaCl, 100 mM KCl, 0.5 mM MgCl2 and 3 mM guanosine-5′-[(β,γ)-methyleno]triphosphate were incubated at room temperature for 10 min. The final concentration of unfiltered Folch liposomes was 0.6 mg ml−1. Samples were incubated on carbon-coated copper grids (Plano) and stained with 2% uranyl acetate.

Yeast growth assay

To test the ability of mutant Mgm1 variants to complement the loss of wild-type Mgm1 in yeast (Saccharomyces cerevisiae), a GAL1 promoter was inserted upstream of the MGM1 open reading frame by homologous recombination. To this end, the GAL1 promoter was amplified from pFA6a-kanMX6-PGAL141 (using oligonucleotides MGM1-PGAL-FW CATCCCAAGAGTGGCGAACTATAACACATTAGTAAGGATGgaattcgagctcgtttaaac and MGM1-PGAL-REV GCTGTCTTCTCAGAATTAAAAGCCGTACTGGGCTCGCATTcattttgagatccgggtttt42) and transformed into the YPH499 wild-type strain43. Mutations were introduced into pRS414-Mgm144 by site-directed mutagenesis. The PGAL1-MGM1 yeast strain was transformed with the empty vector pRS414 or pRS414-Mgm1 encoding wild-type Mgm1 or mutant variants. After selection on synthetic defined (-TRP) media (0.67% (w/v) YNB without amino acids (BD Difco), -TRP amino acid drop-out mix (MP Biomedicals)) containing 2% (w/v) galactose and 1% (w/v) raffinose, yeast were grown in media containing 2% (w/v) glucose as a carbon source to suppress expression of the endogenous wild-type Mgm1 allele. Under these conditions, cells expressing no or non-functional Mgm1 rapidly lose mitochondrial DNA6. Subsequently, cultures were diluted in media containing 0.2% (w/v) glucose in 48-well microtiter plates and growth was monitored for 24 h at 30 °C using a Tecan Spark 10M microplate reader by measuring the absorbance at 600 nm every 5 min after a 10 s linear shake with an amplitude of 2.5 mm at 630 r.p.m. Between cycles, the plate was agitated in a double-orbital manner with an amplitude of 1.5 mm at 180 r.p.m. Blank-corrected mean absorbance values from two or three wells per mutant strain were plotted using GraphPad Prism 6.0 and growth experiments were repeated with cell populations from three independent yeast transformations.

To test for dominant-negative effects of Mgm1 mutants, the wild-type strain YPH499 was transformed with pRS414-Mgm1 encoding wild-type or mutant Mgm1 and growth was assessed in synthetic defined medium containing 3% (v/v) glycerol as described above. To test whether Mgm1 variants are stably expressed in yeast cells and able to retain mitochondrial DNA, mitochondria were isolated on a small scale45 and analysed by SDS–PAGE and western blotting using antibodies directed against Mgm1, Cox1 (mitochondrially encoded cytochrome c oxidase subunit 1) and Ssc1 (mitochondrial Hsp70, loading control).

Yeast microscopy

Yeast cells were grown in synthetic defined (-TRP) media containing either 2% (w/v) glucose (for PGAL1-MGM1 yeast strains expressing plasmid-borne Mgm1 variants) or 3% (v/v) glycerol (for dominant-negative mutant strains) to mid-logarithmic phase and stained with 0.5 µg ml−1 DAPI (4′,6-diamidino-2-phenylindole) and 175 nM DiOC6 (3,3′-dihexyloxacarbocyanine iodide) in 5% (w/v) glucose and 10 mM HEPES (pH 7.2). Immediately after staining, Z-stacks were recorded on a Leica DMi8 fluorescent microscope with a 63×/1.40 objective and a Leica DFC3000 G CCD camera. Images were deconvoluted with Huygens Essential (Scientific Volume Imaging, http://svi.nl) and maximum intensity projections were created in Fiji46. Contrast was adjusted linearly to correct for variations in DiOC6 uptake. For quantification of mitochondrial morphology, cells with tubular or fragmented mitochondrial networks were counted in images from three independent cultures (for each culture at least 70 cells were counted). Wild-type ρ0 cells were generated by ethidium bromide treatment of ρ+ cells.

Electron microscopy of yeast mitochondria

Yeast cells were fixed for 3 h with 4% (w/v) paraformaldehyde and 0.5% (v/v) glutaraldehyde in 0.1 M citrate buffer (pH and temperature adjusted to growth conditions). Samples were treated with 1% (w/v) sodium metaperiodate for 1 h at room temperature. Yeast cells were embedded in 10% (w/v) gelatin, infiltrated with 2.3 M sucrose and frozen in liquid nitrogen. Ultrathin sections were cut at −115 °C (Reichert Ultracut S, Leica) and collected on 200-mesh copper grids (Plano) coated with Formvar and carbon. Sections were stained with 3% (w/v) tungstosilicic acid hydrate in 2.5% (w/v) polyvinyl alcohol. Samples were examined at 80 kV with a Zeiss EM 910 electron microscope (Zeiss), and images were recorded with a Quemesa CCD camera and the iTEM software (Emsis). Images were analysed by ImageJ/Fiji46. All applied statistical tests were calculated using Prism (GraphPad software). A normality distribution test (Kolmogorov–Smirnov test) was carried out for all experimental values, and with normally distributed data a Student’s t-test (two-tailed P value) was applied, otherwise the Mann–Whitney rank-sum (two-tailed P value) test was used to calculate the significant difference between two groups.

Liposome preparation for cryo-ET

For examining Mgm1 assembly on membranes by electron cryo-tomography, dried lipids were rehydrated to a final concentration of 3 mg ml−1 in liposome buffer (20 mM HEPES, pH 7.5, 150 mM NaCl). Folch lipids (brain extract from bovine brain, type I, fraction I, Sigma-Aldrich) were used for inside decoration, or a lipid mixture of 70% galactocerebroside47, 10% cardiolipin (both Sigma-Aldrich) and 20% di-oleyl-phosphatidylcholine (DOPC) (Avanti Polar Lipids) for outside decoration of tubes. Liposomes were prepared by sonication followed by extrusion through a 1 μm polycarbonate filter. Rehydrated lipids were incubated with purified Mgm1 (final concentration 10 μM) for 30 min at room temperature in the absence or presence of GTPγS (final concentration 1 mM, Jena Bioscience). For inside decoration, Mgm1 (with or without nucleotide) was added before the liposome preparation step.

Grid preparation and image acquisition for electron cryo-tomography

The final sample was mixed in a 1:1 ratio with colloidal gold fiducial markers and 3 μl were applied to freshly glow-discharged R2/2 Cu 300-mesh holey carbon-coated support grids (Quantifoil Micro Tools). Grids were plunge-frozen using a Vitrobot Mark IV plunge-freezer at 100% humidity and 10 °C. Samples were imaged in a FEI Titan Krios electron microscope (FEI Company) operating at 300 kV, equipped with a K2 summit direct electron detector and Quantum energy filter (Gatan). The nominal magnification was set to 53,000×, yielding a calibrated pixel size of 2.7 Å. Tomographic tilt series were acquired following a dose-symmetric tilting scheme48 with a 3° increment and a cumulative total electron dose of approximately 90 e Å−2. Defocus values ranged from −2.0 to −4.0 μm. Data were acquired with the SerialEM software package49 in dose-fractionation mode.

Tomogram reconstruction and subtomogram averaging

Dose-fractionated movies of tomograms were aligned using either Unblur50 or MotionCor251. After contrast-transfer-function correction, images were combined to generate a raw image stack that was used as input for generating tomograms with IMOD. Single tilt-images were aligned by gold fiducial markers and volumes reconstructed by weighted back-projection. Particle extraction, alignment and subtomogram averaging were performed with Dynamo52 and MATLAB. For a whole tube, particles were picked along the filaments using the respective option in the Dynamo toolbox. Eighteen membrane tubes covered with a clear visible protein coat within 15 different tomograms and 10 tubes within 10 tomograms were used for processing for the apo form and the GTPγS bound form, respectively. Owing to the differences in diameter of the inside decoration, only two membrane tubes in two individual tomograms were used for the apo form as well as the GTPγS bound state. For close-up views, tubes were sub-boxed along the helical pattern. For tubes decorated on the inside, particles were picked along the wall of the lipid tube. Before subtomogram averaging, the datasets were divided into two independent half sets for resolution estimation. Each half set was aligned to an independent reference generated from a subset of each half set and reference-free alignment. To address the possibility of different handedness, classification was performed during the processing workflow. Only protein assemblies with a left-handed helical pattern were observed. To exclude that the left-handed arrangement of the outside decoration was driven by the preformed lipid tubes, subtomogram averaging of Mgm1 covering the outside of Folch lipid tubes of different diameters was performed. Also in this case, only protein assemblies with a left-handed helical pattern were observed. The numbers of particles that contributed to the converged averages of the main structures and final resolution from Fourier shell correlation (FSC) curves are listed in Extended Data Table 2. The final structures were obtained using relion_reconstruct from the Relion toolbox. USCF Chimera and MATLAB were used for structure and FSC curve display, respectively53.

Molecular dynamics simulations

Flexible fitting into cryo-ET volume

A general approach for building atomic models from cryo-ET reconstructions is to include a potential energy term coupling the atomic coordinates during a molecular simulation to the experimentally determined density. Here we used the MDfit method54, which uses an all-atom structure-based model55 based on the tetramer crystal structure, and additionally includes an energetic term that attempts to maximize the correlation between the experimental density and the simulated density of the molecular dynamics trajectory. The structure-based model has an explicit energy minimum at the tetramer crystal structure, which means that the secondary structure seen in the crystal is maintained during flexible fitting. Modified Gromacs source code containing MDfit and software for creating the all-atom structure-based topologies are available for download at55 http://smog-server.org. Default MDfit parameters were used, including setting the energetic weight of the map equal to the number of atoms. For both the inner and outer decoration, the initial configuration was generated by manually placing twelve tetramers (247,728 heavy atoms) into and surrounding the cryo-ET volume with the aid of the “Fit in Map” tool in Chimera53. Simulations were performed until the cross-correlation stabilized. Only the dimers that were completely within the cryo-ET volume were saved for deposition alongside the cryo-ET volume. After fitting the inner decoration, G domains appeared to be in contact. This was checked by strongly constraining the G domains to form the G interface (as in dynamin). The fit including the constraint was nearly identical to that without, suggesting that the cryo-ET for the inner decoration contains the canonical G interface. The submitted model includes the constraint.

All-atom molecular dynamics to support the “pre-shaped” tetramer and characterize its flexibility

A 2.6-μs all-atom molecular dynamics simulation of a stalk tetramer in explicit solvent was performed to estimate its shape in the absence of crystal interactions. The simulation was initialized from the tetramer crystal structure with a closed interface-1 and contained for each monomer the four stalk helices (residues 549–590, 635–720 and 828–877). Two G–G–S–G–G linkers were used to connect breaks in the stalk where the paddle was cut out, creating a single chain for each stalk monomer. Simulations were performed with Acellera ACEMD56 using the CHARMM36 forcefield57. Details of the simulation are as follows: NPT ensemble, temperature 300K, Langevin thermostat, Berendsen barostat at 1 atm, restrained bonds, timestep 4 fs, PME electrostatics, grid spacing 1 Å, cutoff 9 Å, switching at 7.5 Å. The conformation of the stalk tetramer was analysed to estimate the structural preference and flexibility of a stalk filament containing the tetramer. See Extended Data Fig. 7 for details.

All-atom structure-based model for inner decoration of 1-start helix

Our aim was to determine the tetramer structure upon confinement in a filament decorating the interior of a narrow membrane tube (r = 30 nm) with a small pitch (P = 12 nm). In particular, we were interested in whether the crystallographic interfaces-1 and -2 can be consistent with negatively curved geometries. To this end, a molecular dynamics simulation was performed on a short filament (octamer) using a simplified potential that includes the all-atom geometry. Three constraints were imposed: the putative membrane-binding residues R748 and K749 in each monomer were constrained to a 30 nm radius from the z-axis; an impenetrable cylindrical wall was imposed with a 30 nm radius; and the z coordinate of the centres of mass of each dimer (interface-2) were constrained such that the short filament had an effective pitch of 12 nm. No restraints were introduced in interface-1. The simulation potential was an all-atom structure-based model using the tetramer crystal structure with interface-1 formed. The simulation topology for Gromacs58 was created using the tetramer crystal and SMOG2.1 with the default forcefield ‘SBM_AA’55. The octamer topology was created by merging two tetramer topologies and additionally copying the requisite pair interactions for the new interface-1 created by connecting the tetramers. Langevin dynamics with a low temperature (0.16 reduced units, 20K Gromacs temperature) for 10 × 106 steps was used to get near to the minimum energy subject to the constraints. A steepest-descent minimization was used for the final analysed configuration. To minimize edge effects, the interior tetramer of the octamer filament was analysed.

Tube-pulling assays

Mgm1 was labelled with a fluorescein-labelled peptide using a sortase-mediated reaction59. All lipids were purchased from Avanti Polar Lipids. GUVs were electroformed24 from a lipid mix (2 mg ml−1) containing DOPC, di-oleyl-phosphatidylserine, rhodamine-phosphatidylethanolamine (Rhod-PE) and di-sialyl-phosphatidylserine-polyethylene-glycol-2000-biotin (DSPE-PEG(2000)Biotin), at a ratio of 7:3:0.01:0.003. GUVs were then transferred to a microscopy chamber of two rectangular glass slides (11 × 35 mm) and mounted on an inverted microscope including a Nikon Eclipse Ti base, a CSU-X1 confocal system (Nikon), an Andor Ixon EMCCD camera (Oxford Instruments) and homemade optical tweezers consisting of a 5-W, 1,064-nm laser (ML5-CW-P-TKS-OTS, Manlight) focused through a 100 × 1.3 numerical aperture oil objective. Images were acquired using SlideBook software (Intelligent Imaging Innovation). Bead traces were acquired with a C-MOS Camera (Picelink) using custom-made software. Outward membrane nanotubes were formed by holding a 3.05-µm streptavidin-coated polystyrene bead (Spherotech) glued onto a GUVs with optical tweezers, while pulling away the GUVs held by aspiration with a hand micropipette and controlled with motorized micromanipulators (MP-285, Sutter Instrument). Subsequently, Mgm1 was diluted to a final concentration of 3 μM in 20 mM HEPES/NaOH pH 7.4, 200 mM NaCl and 1 mM MgCl2, and injected in the vicinity of the membrane tube using a second micropipette connected to a pressure control system (MFCS-VAC, −69 mbar, Fluigent). For pulling membrane nanotubes inward, 2.01 µm glass beads (Bangs Laboratories) were internalized with optical tweezers into GUVs adhering to an Avidin-coated flow chamber (coverslip and sticky-Slide VI 0.4, Ibidi). Tubes were pulled by moving the stage, and thus the GUV. Mgm1 (3 μM) was added with a syringe pump (Aladdin, World Precision Instruments) connected to the Ibidi flow chamber. The force F was determined by applying Hooke’s law F = kΔx to the bead displacement Δx and trap stiffness k (3.05 µm beads: k = 79 pN nm−1; 2.01 µm beads: k = 75 pN nm−1). The basis of inward-pulled tubes was unstable and moved on the surface of the GUV, so that the projection on the bead displacement in the x and y axes changed rapidly. Furthermore, because the beads were pre-endocytosed into the GUVs, the initial position of the bead without force was unknown, as compared with the outward-tube pulling assay in which the bead position was recorded before it became attached to the GUV. Therefore, ΔF instead of F was plotted as it is more reliable. In Extended Data Fig. 8d, 6 μM Mgm1 was added to increase protein polymerization and therefore the force generated. In experiments requiring GTP, the buffer was supplemented with 2 mM GTP. The following settings were applied for Fig. 4d, Extended Data Fig. 8a, b: resolution: 512 × 512 × 10 s, 145 nm per pixel, 16 bit; fluorochromes: fluorescein (excitation: 488 nm, bandpass filter 520/50, dichroic beamsplitter 405/488/568/647; LUT: green (Fiji)) rhodamine B (excitation: 561 nm, bandpass filter 607/30, dichroic beamsplitter 405/488/568/647; LUT: red (Fiji)); experiments were performed at room temperature in 20 mM HEPES/NaOH pH 7.4, 200 mM NaCl and 1 mM MgCl2. For Fig. 5c, Extended Data Fig. 8c, d: resolution: 512 × 512 × 30 s, 145 nm per pixel, 16 bit; fluorochromes and experimental conditions as above.

Reporting summary

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