Main

We identified apcin (Fig. 1a) in an earlier study as an inhibitor of cyclin proteolysis in mitotic Xenopus egg extract17, but its mechanism of action remained unknown. Analysis of the structure–activity relationship (Fig. 1b and Extended Data Fig. 1a) revealed that elimination of one nitrogen in the pyrimidine ring of apcin (apcin-P) reduced activity slightly, but replacement with a morpholino group (apcin-M) eliminated activity. In contrast, elimination of the nitro-imidazole moiety (apcin-A) had little effect. To identify the target of apcin, we coupled apcin-A to beads via its amino group, incubated the beads with mitotic Xenopus extract, then removed the beads. We found that Cdc20 was depleted from the extract (Fig. 1c), resulting in stabilization of a cyclin-B-luciferase reporter protein (Fig. 1d). Cyclin degradation was rescued by adding in vitro-translated Cdc20 (Fig. 1d), implicating Cdc20 as the target of apcin. Cdc20 binding to apcin-A beads could be competed by free apcin (Fig. 1e), but not by the inactive analogue apcin-M or the APC/C inhibitor tosyl-l-arginine methyl ester (TAME)15,16 (Extended Data Fig. 1b). A cyclin B1 amino (N)-terminal fragment (cycB1-NT) also competed for Cdc20 binding to the apcin-A resin, but the same fragment with a mutated D-box did not (Fig. 1f). Among a panel of WD40-containing proteins, Cdc20 binding to apcin-A beads was most robust, followed by Cdh1, with much less binding of other WD40-containing proteins observed (Extended Data Fig. 1c). Although we have not tested the ability of apcin to inhibit Cdh1-dependent ubiquitylation, apcin inhibited Cdh1-dependent proteolysis in interphase Xenopus extract less efficiently than Cdc20-dependent proteolysis in mitotic Xenopus extract (Extended Data Fig. 1d). Apcin bound to endogenous Cdc20 in Xenopus extract in a dose-dependent manner (Extended Data Fig. 1e) that correlated with its ability to inhibit formation of high-molecular-mass ubiquitin conjugates of cycB1-NT (Fig. 1g) or full-length cyclin B1 (Extended Data Fig. 2a). Kinetic analysis of a reconstituted APC/C-dependent ubiquitylation reaction16 showed that apcin caused a significant increase (P = 0.0039) in the Michaelis constant, Km (inhibition constant Ki = 23 µM), but no reduction in the catalytic rate constant kcat (Extended Data Fig. 2b). Together these results suggest that apcin competitively inhibits APC/C-dependent ubiquitylation by binding to Cdc20 and preventing substrate recognition.

Figure 1: Apcin binds to Cdc20 and inhibits APC/C-dependent ubiquitylation.
figure 1

a, Structures of apcin and derivatives. b, Effects of apcin and derivatives (200 μM) on proteolysis of an N-terminal fragment of cyclin B1 (cycB1-NT) in mitotic Xenopus egg extract. Substrate levels were measured at 40 min. N = 3 independent experiments. c, Apcin-A resin depletes Cdc20 from mitotic Xenopus egg extract. Cdc20 and Cdc27 levels were measured by western blotting. d, Depletion with Apcin-A resin stabilizes a cyclin-luciferase reporter protein and degradation can be rescued by addition of in vitro-translated Cdc20. Substrate levels were measured at 60 min. N = 3 independent experiments. e, Cdc20 expressed in reticulocyte lysate binds to apcin-A resin and can be competed by free apcin. Cdc20 was detected by western blotting. N = 3 independent experiments. f, Wild-type (WT) cycB1-NT, but not a D-box mutant (DBM, mutation of R42A, L45A), competes with Cdc20 binding to apcin-A resin. Cdc20 was detected by western blotting. N = 4 independent experiments. g, Apcin inhibits formation of high-molecular-mass ubiquitin conjugates of [35S]cycB1-NT in mitotic Xenopus extract. Bars, mean ± s.e.m.

PowerPoint slide

To identify the site on Cdc20 that binds apcin, we soaked apcin into Cdc20 protein crystals and determined the structure of the Cdc20–apcin complex to 2.1 Å resolution (Extended Data Table 1). We found that apcin bound a small pocket on the side of the WD40 domain that has been implicated in binding the D-box (Fig. 2a, b and Extended Data Fig. 3a, b)14,18,19. The binding mode of apcin is consistent with the structure–activity relationship, as the pyrimidine ring and aminal nitrogens make hydrogen bonds with backbone atoms from D177. The hydrophobic trichloromethyl group is buried in the pocket occupied by leucine of the D-box (Extended Data Fig. 3b). The nitro-imidazole moiety is positioned facing solvent, explaining why apcin-A retains activity and can be used to isolate Cdc20 when coupled to beads.

Figure 2: Apcin binds to the D-box binding site of Cdc20.
figure 2

a, Crystal structure of the apcin–Cdc20 complex. Apcin atoms are labelled in yellow (carbon), blue (nitrogen), red (oxygen) and green (chlorine). Cdc20 is shown in magenta. Dotted blue lines indicate hydrogen bonds. b, View is rotated to show the position of V200 at the base of the hydrophobic binding pocket. c, Mutation of residues in the binding pocket reduces Cdc20 binding to apcin-A resin (red bars) and the capacity of in vitro-translated Cdc20 protein to rescue cyclin-luciferase degradation in mitotic Xenopus egg extract immunodepleted of Cdc20 (blue bars). Bars, mean ± s.e.m. of three independent experiments. P values were calculated by an unpaired t-test compared with WT.

PowerPoint slide

We tested whether mutations in the presumptive binding site affect the binding of Cdc20 to apcin-A beads (Fig. 2c and Extended Data Fig. 3c, d). V200 is located at the base of the hydrophobic pocket and mutation to methionine blocked binding of Cdc20 to apcin-A beads and inactivated the ability of Cdc20 to rescue cyclin degradation in a Cdc20-depleted extract. Mutation of D177, P179 or I216, which line the binding pocket, also reduced binding to apcin and function of Cdc20. Mutation of R174, which lies near the pyrimidine ring of apcin, also reduced apcin binding and blocked Cdc20 rescue activity, consistent with a role of this residue in interacting with negatively charged amino acids in the D-box14. In contrast, mutation of E465, which interacts with the conserved arginine of the D-box14 but lies distant from the apcin-binding pocket, decreased the ability of Cdc20 to rescue degradation, but had little effect on apcin binding. E180 lies further away and mutation to alanine had no effect on apcin binding or rescue activity. Overall, we observed a strong correlation between effects of mutations on apcin binding and their effects on Cdc20 function for residues that line the apcin-binding pocket.

Cdc20 is recruited to the APC/C through multiple weak interactions6,8,20,21,22,23. For example, substrates can promote cooperative Cdc20 binding to the APC/C through a co-receptor interaction in which the substrate is simultaneously recognized by Cdc20 and the APC/C6,7,8,9,10,11,12,13,14. Consistent with this idea, we found that addition of substrate increased Cdc20 loading onto the APC/C in a concentration- and D-box-dependent manner in Xenopus extract (Fig. 3a and Extended Data Fig. 4a, b). Substrate-induced loading of Cdc20 could be blocked by addition of apcin (Fig. 3a and Extended Data Fig. 4a), indicating that the leucine-binding pocket of Cdc20 is critical for co-receptor function. Because the binding of Cdc20 to the APC/C was variable in the absence of added substrate, we were not able to assess the effects of apcin under this condition. Thus we cannot exclude the possibility that apcin might also decrease Cdc20 binding to the APC/C in the absence of substrate. The small molecule TAME, which antagonizes the Ile-Arg (IR)-tail interaction between Cdc20 and the APC/C15,16, also antagonized Cdc20 loading. At high concentrations of substrate, the combined use of apcin and TAME was more effective at blocking Cdc20 binding to the APC/C than either compound used alone, suggesting that simultaneous disruption of multiple interactions between substrate, Cdc20 and APC/C may be an effective strategy for inhibiting APC/C.

Figure 3: Effects of apcin on Cdc20 binding to APC/C and stability of APC/C substrates in mitotic Xenopus extract.
figure 3

a, Apcin blocks co-receptor-dependent binding of Cdc20 to APC/C. Substrate (cycB1-NT, 1 μM or 10 μM), apcin and/or TAME (50 μM each) were added, and APC/C was isolated with anti-Cdc27 antibody. Levels of Cdc20 were assessed by western blotting and normalized to the levels of Cdc27. b, Effect of apcin and TAME on APC/C substrate stability. Levels of 35S-labelled substrates were assessed by gel electrophoresis and phosphorimaging. c, As in b except that combinations of apcin and TAME were examined. Values and error bars in ac represent means ± s.e.m. of three independent experiments.

PowerPoint slide

Because substrates can be recruited to the APC/C through both D-box-dependent and -independent mechanisms, we compared the ability of apcin to stabilize different APC/C substrates in mitotic Xenopus extract. Apcin stabilized cycB1-NT and securin most effectively (Fig. 3b and Extended Data Figs 4c and 5b), with somewhat weaker effects against full-length cyclin B1 (Fig. 3b and Extended Data Fig. 4c). Interestingly, even high concentrations of apcin failed to stabilize cyclin A2 or Nek2A (Fig. 3b and Extended Data Figs 4c and 5a), suggesting that apcin is highly specific for the D-box binding site. In contrast, TAME inhibited the degradation of all APC/C substrates tested (Fig. 3b and Extended Data Figs 4c and 5a, b), consistent with its ability to block recruitment of Cdc20 to the APC/C directly15,16. The ability of apcin to inhibit formation of high-molecular-mass substrate–ubiquitin conjugates closely correlated with its ability to stabilize each substrate (Extended Data Fig. 6). These results suggest that the leucine-binding pocket of Cdc20 plays an essential role in recruiting the D-boxes of securin and cycB1-NT, and that apcin can effectively compete with these interactions. The ability of full-length cyclin B1 and cyclin A2 to bind Cks1 via Cdk1 may facilitate their D-box-independent recruitment to the APC/C24,25, helping these substrates partially overcome the effects of apcin. Furthermore, the N-terminal region of cyclin A2 appears to bind Cdc20 with higher affinity than cyclin B1 (ref. 26), which may further reduce the effectiveness of apcin. In contrast to these substrates whose degradation is D-box-dependent, Nek2A is recruited directly to the APC/C via a Cdc20-independent mechanism that requires its Met-Arg (MR)-tail rather than a D-box27,28,29, explaining why apcin fails to inhibit ubiquitylation or degradation of this protein.

Because apcin and TAME inhibit APC/C-dependent proteolysis by distinct mechanisms, we tested the effect of combining the inhibitors on the proteolysis of APC/C substrates in Xenopus extract. The combination of apcin and TAME led to synergistic stabilization of cyclin B1, cycB1-NT, securin and cyclin A2, with a much weaker effect for Nek2A (Fig. 3c and Extended Data Figs 4d and 5a, b). For example, combining TAME and apcin at 25 µM each was more effective at stabilizing cyclin B1 than using either compound alone at 100 µM. Apcin slightly enhanced the ability of TAME to stabilize Nek2A (Fig. 3c), suggesting that the leucine-binding pocket of Cdc20 may bind Nek2A, even though this interaction is not essential for proteolysis if the APC/C is not otherwise perturbed.

We next examined the effect of apcin, proTAME15 (a cell-permeable TAME prodrug), and the combination, on mitotic exit in four different human cell lines. Apcin and proTAME synergized to increase the mitotic fraction in all cell lines examined (Fig. 4a and Extended Data Fig. 7a, b). Apcin-M was inactive whereas apcin-P retained activity (Extended Data Fig. 7c), consistent with effects on Cdc20 binding and APC/C-dependent proteolysis in Xenopus extract. In live-cell imaging experiments in RPE1 cells (Fig. 4b, Extended Data Fig. 8a–d and Supplementary Videos 1 and 2), 25 µM apcin had no detectable effect on mitotic duration (P = 0.279) unless mitosis was artificially shortened by depletion of the spindle assembly checkpoint (SAC) protein Mad2 (P = 0.0001). In contrast, in the presence of proTAME, the addition of apcin dramatically slowed the rate of mitotic exit in a synergistic manner (Fig. 4b and Extended Data Fig. 8a–d): the rate of mitotic exit was 63% of that predicted by a multiplicative combination of the single compound effects (P = 0.016). Significant synergy was also observed in U2OS cells (P = 2.0 × 10−8; Extended Data Fig. 8e). Addition of the inactive derivative apcin-M had no effect on the rate of mitotic exit in the presence of proTAME (P = 0.68; Fig. 4b and Extended Data Fig. 8a–e). The response of cells to proTAME alone was biphasic, because prolongation of metaphase can cause cohesion fatigue30, which reactivates the SAC to block mitotic exit in a subpopulation of cells. Notably, the addition of apcin eliminated the biphasic response. Furthermore, the combined effect of apcin and proTAME was largely preserved when the SAC was inactivated by Mad2 depletion. When modelled quantitatively, the degree of synergy between apcin and proTAME was in fact enhanced in the absence of Mad2, as the rate of mitotic exit was reduced to 38% of the rate predicted by a multiplicative combination of the single compound effects (P = 8.92 × 10−5; Extended Data Fig. 8d). Together these findings suggest that synergistic inhibition of mitotic exit does not rely on the SAC, but instead probably arises from direct pharmacological APC/C inhibition.

Figure 4: Apcin synergizes with proTAME to prolong mitotic duration.
figure 4

a, RPE1 cells were treated with indicated concentrations of apcin and proTAME for 18 h, fixed and the mitotic index determined by automated high-throughput imaging. The panel shows the difference between the mitotic indices calculated by a Bliss-independence model compared with a synergy model; any positive value indicates synergy. *P < 0.05 on the basis of analysis of four technical replicates. b, Asynchronous RPE1 cells expressing H2B–green fluorescent protein (GFP) were transfected with short interfering RNA (siRNA) and 24 h later treated with apcin or apcin-M (25 μM) and/or proTAME (6 μM). Cells were then imaged every 6 min for 45 h. Mitotic duration and cell fate (Extended Data Fig. 8b) were determined by manual inspection of the videos and plotted as inverse cumulative frequency (−proTAME) or Kaplan–Meier curves (+proTAME). The hatch marks on the Kaplan–Meier curves indicate censored cells that did not exit mitosis or die in mitosis before the end of the movie or before they migrated out of the field of view. Graphs include the combined results of two independent experiments.

PowerPoint slide

Source data

The discovery of apcin and the elucidation of its mechanism validate the importance of the leucine-binding pocket in Cdc20 as a D-box co-receptor. Apcin preferentially stabilizes APC/C substrates whose degradation is D-box dependent. However, on its own, low concentrations of apcin do not effectively block mitotic exit, either because substrates can outcompete apcin binding to the leucine-binding pocket, or because substrates can be recruited to the APC/C through other mechanisms. The effectiveness of apcin can be dramatically enhanced by the addition of TAME, which blocks Cdc20 loading through a distinct mechanism, highlighting the importance of multiple weak protein–protein interactions in promoting activator binding and efficient substrate ubiquitylation (Extended Data Fig. 9). Our work highlights the possibility of disrupting the function of a protein machine by simultaneously inhibiting multiple protein–protein interactions. Because dynamic protein complexes regulate virtually all aspects of cell biology, simultaneous targeting of multiple weak interactions may represent a new opportunity for therapeutic targeting of protein complexes that may otherwise be difficult to inhibit with a single compound.

Methods

Reagents

Commercial antibodies used for western analysis were as follows: anti-Cdc27 (610455, BD Transduction Laboratories), anti-Cdc20 (BA8; sc-93399, Santa Cruz Biotechnology and NB 100-2646, Novus Biologicals) to recognize Xenopus Cdc20, anti-Cdc20 (H-175; sc-8358, Santa Cruz Biotechnology) to recognize human Cdc20, and anti-HA-peroxidase (3F10; 12013819001, Roche). Secondary antibodies used included anti-rabbit IgG-HRP (NA934; GE Healthcare) and anti-mouse IgG-HRP (NA931; GE Healthcare). For APC/C immunopurification from Xenopus extract, anti-Cdc27 (AF3.1; sc-9972, Santa Cruz Biotechnology) was used. For immunodepletion of Xenopus Cdc20, a rabbit polyclonal antibody was generated by Yenzym by immunization with an N-terminal fragment of Cdc20 (residues 1–170; tagged at the carboxy (C) terminus). Chemicals used were cycloheximide (239764, Calbiochem), calcium ionophore A23187, free acid form (100105, Calbiochem), tosyl-l-arginine methyl ester (T4626, Sigma), proTAME (I-440, Boston Biochem), MG262 (I-120, Boston Biochem), apcin (T0506-3874, Enamine), apcin-P (Amb2237944, Ambinter) and apcin-M (Amb1395012, Ambinter). Apcin and apcin-A were also synthesized by Sundia Meditech according to the methods described in the Supplementary Information.

Assessment of substrate degradation in Xenopus egg extract

Use of female Xenopus laevis to produce cytoplasmic egg extracts was approved by the Harvard Medical School Standing Committee on Animals (protocol number 03231). Interphase Xenopus egg extract was prepared from eggs laid overnight according to the protocol of ref. 31 with the exception that eggs were activated with 2 μg ml−1 calcium ionophore (A23187) for 30 min before the crushing spin. Extract was frozen in liquid nitrogen and stored at −80 °C. Interphase extract was induced to enter mitosis by addition of non-degradable cyclin B (MBP-Δ90) at 20 μg ml−1 and incubated at 22–24 °C for 30–60 min. MBP-Δ90 consists of a fusion of the maltose-binding protein (MBP) to Xenopus cyclin B1 lacking its N-terminal 90 amino acids32 and was expressed in Escherichia coli by inducing cultures at an attenuance (D600 nm) of 0.6 with 300 μM isopropylthiogalactoside (IPTG) for 5 h at 22–24 °C. Purification followed New England BioLabs protocol. To promote degradation in interphase extract, Cdh1 protein, expressed in baculovirus, was added to extract at a final concentration of 50 nM in the presence of 75 μM roscovitine. Roscovitine addition is necessary to suppress inhibitory phosphorylation of Cdh1 by Cdk1. Extract was then pre-treated with drug (dimethylsulphoxide (DMSO), apcin and/or TAME) for 15 min at 22 °C before addition of substrates.

Substrates consisted of human full-length cyclin B1, cyclin A2, securin, Nek2A or an N-terminal fragment (residues 1–88) of human cyclin B1 (cyc B1-NT). Each substrate was amplified with primers by PCR to allow T7-dependent transcription of the PCR product. Substrates were expressed and labelled with [35S]methionine (Perkin Elmer NEG709A500UC) using the TNT system (Promega). To measure degradation of substrates, extract was pre-treated with DMSO or test compounds for 10 min in the presence of 100 μg ml−1 cycloheximide to prevent re-incorporation of free labelled amino acid. The in vitro translation reaction was then added to the Xenopus extract at 10% final volume. Extract was then incubated at 24 °C, with shaking at 1,250 r.p.m., with samples taken at indicated times. Reactions were quenched with sodium dodecyl sulphate (SDS) sample buffer and processed for SDS–polyacrylamide gel electrophoresis (PAGE) and phosphor imaging (Bio-Rad PMI); quantification was performed using Quantity One software (Bio-Rad).

Assessment of substrate ubiquitylation in mitotic Xenopus egg extract

Interphase Xenopus egg extract was supplemented with MBP-Δ90 to promote entry into mitosis. Mitotic extract was then treated for 30 min at 24 °C with 20 μM ubiquitin-vinyl sulfone (UbVS; U-202, Boston Biochem) to suppress deubiquitylation and with 100 μg ml−1 cycloheximide to prevent re-incorporation of free labelled amino acid. Subsequently, mixtures containing apcin or DMSO, as indicated, proteasome inhibitor MG262 (150 μM) and wild-type ubiquitin (44 μM) were added to UbVS-treated extract and incubated for additional 10 min, at 24 °C with agitation. In vitro translation reactions expressing human full length cyclin B1, cyclin A2, securin, Nek2A, or cycB1-NT with [35S]methionine labelling were also pre-treated with UbVS at 20 μM and added to pre-treated extract at 10% final volume. Extract was then incubated at 24 °C, with shaking at 1250 r.p.m., and substrate ubiquitylation monitored by taking samples at indicated times. Reactions were quenched with SDS sample buffer and processed for SDS–PAGE and phosphor imaging (Bio-Rad PMI), and quantification was performed using Quantity One software (Bio-Rad).

Measurement of the apcin-Cdc20 interaction in Xenopus egg extract using cellular thermal shift assay

The cellular thermal shift assay (CETSA) method33 was adapted to examine apcin engagement of endogenous Cdc20 in Xenopus egg extract. Interphase extract was diluted tenfold and incubated with various concentrations of apcin dissolved in DMSO in a total volume of 200 μl, with a final DMSO concentration of 1%. After 10-min incubation at 22–24 °C, 50 μl of the lysate was transferred into PCR tubes (20170-012, VWR) and heated in a PCR machine (Mastercycler gradient, Eppendorf) for 3 min at 46 °C, followed by cooling for 3 min at 22–24 °C. These conditions were established in preliminary experiments as the temperature that yielded the greatest degree of Cdc20 precipitation that could be rescued by apcin treatment (data not shown). The heated lysates were centrifuged at 14,000 r.p.m. (20,000g) for 20 min at 4 °C to separate the soluble fractions from precipitates. Twenty microlitres of the supernatants were mixed with SDS sample buffer and the fraction of soluble Cdc20 was analysed by SDS–PAGE and anti-Cdc20 western blot. Quantification of soluble Cdc20 used Fuji Imager LAS3000 and ImageJ software. Soluble Cdc20 levels were normalized to soluble Cdc20 in samples treated with the highest concentration of apcin.

Measurement of Cdc20 binding to APC/C in Xenopus egg extract

To examine levels of Cdc20 associated with APC/C, the APC/C was immunopurified from mitotic Xenopus egg extract. For 100 μl extract, 2 μg of anti-Cdc27 antibody (AF3.1, Santa Cruz Biotechnology) was cross-linked to 5 μl of Affiprep Protein A beads (156-0006, Bio-Rad) and incubated for 1 h at 4 °C. Apcin, TAME or DMSO was mixed with extract upon addition to anti-Cdc27-Affiprep Protein A beads in the presence or absence of cycB1-NT containing a HA-tag at the N terminus and His tag at the C terminus, as previously described34. After incubation with extract, beads were washed quickly three times with 20-fold volume of XB (10 mM potassium HEPES, pH 7.7, 100 mM KCl, 0.1 mM CaCl2, 1 mM MgCl2) and combined with SDS sample buffer. For analysis of Cdc20 binding to APC/C, samples were processed for SDS–PAGE and immunoblotting against Cdc20 and APC/C subunit Cdc27. Chemiluminescence was imaged on a Fuji LAS 3000 with Image Reader LAS-3000 software. Levels of Cdc20 were quantified using ImageJ and data normalized to respective Cdc27 levels.

Coupling of apcin-A to affigel-10 resin

Affigel-10 resin (153-6099, Bio-Rad) was washed twice with DMSO and dried. The resin was then mixed with 5 mM or 15 mM apcin-A dissolved in DMSO (2× volume of dry resin). N,N-diisopropylethylamine was diluted 50-fold into the solution. The resin was rotated at 22–24 °C for 2 h and the reaction was quenched with 1/5 resin volume of ethanolamine. The resin was then washed sequentially with isopropanol, water and XB + 0.05% Tween. The resin was stored at 4 °C as a 50% slurry in XB + 0.05% Tween.

Cdc20 depletion by apcin-A resin

For a round of depletion of Cdc20 from mitotic Xenopus extract, apcin-A resin (15 mM coupling) was incubated with extract at 4 °C rotating for 30 min. The volume of resin used was 40% of the extract volume. Three rounds of depletion were performed. To rescue degradation in the depleted extract, reticulocyte lysate containing in vitro-translated human Cdc20 or control reticulocyte lysate, treated as for a translation reaction but with no DNA template, was added to the extract at 1/10th extract volume.

Assay for Cdc20 binding to apcin-A resin

Human Cdc20 in pCS2 vector was mutated at the various residues described with the QuikChange Site-Directed Mutagenesis Kit (Agilent Technologies) and custom primers for each sequence. All mutations were confirmed by sequencing. For a single pull-down assay, 5 μl of apcin-A resin (5 mM coupling) was incubated with 30 μl diluted in vitro-translated human Cdc20 or other WD40 proteins (5 μl reticulocyte lysate diluted to 30 μl with XB + 0.05% Tween) at 24 °C with shaking (1500 r.p.m.) for 30 min. Competitors were pre-incubated with Cdc20 in reticulocyte lysate for 2 min at 22–24 °C before adding to apcin-A resin. Bound and input Cdc20 were detected by western blotting (Fig. 1e, f). Cdc20, other WD40 proteins, ODC and Cdc20 mutants in Fig. 2c and Extended Data Figs 1b, c and 3d were labelled with [35S]methionine and detected by phosphorimaging.

Assessment of kinetics of APC/C-dependent ubiquitylation in a reconstituted system

Measurements of kinetics of ubiquitylation used cycB1-NT and APC/C isolated from Xenopus extract, exactly as previously described16, in the presence or absence of 50 μM apcin. The Ki was calculated on the basis of the assumption of a competitive inhibition model according to the equation Km, apcin = Km, untreated (1 + [I]/Ki), where [I] = 50 μM; Km, apcin (773 nM) and Km, untreated (245 nM) were the average values from three independent experiments. The P value was calculated by an unpaired t-test.

Antibody-based depletion of Cdc20 from Xenopus extract

Cdc20 antibody, covalently coupled to Affiprep protein-A beads as described15, was incubated with mitotic extract at 4 °C with rotation for 30 min. The volume of antibody beads used was 20% of the reaction volume. Three rounds of depletion were performed, with separation of extract from beads after each round by centrifugation in spin columns (89868, Thermo-Pierce).

Luciferase assay

A fusion of the N-terminal domain of cyclin B1 to luciferase17 was added to mitotic extract at 4 μg ml−1 (Fig. 1d) or to interphase extract at 250 μg ml−1 for 10 min at 22–24 °C then diluted to a final concentration of 4 μg ml−1 in mitotic extract (Fig. 2c). The extract was incubated at 22–24 °C and 3 μl samples were taken at 0, 20, 40 and 60 min. The samples were mixed quickly with 30 μl luciferin assay buffer (270 µM coenzyme A, 20 mM tricine, 3.67 mM MgSO4, 0.1 mM ethylediaminetetraacetic acid (EDTA), 33.3 mM dithiothreitol (DTT), 530 µM ATP and 470 µM luciferin, pH 7.8), and the level of luminescence was measured on a Wallac 1420 multilabel counter.

Protein purification and crystallization

The coding region of human Cdc20 containing residues 161–477 (Cdc20–WD40) was amplified by PCR and cloned into the modified pFastBac vector. Recombinant baculovirus encoding the N-terminal His6-tagged Cdc20 protein was constructed using the Bac-to-Bac system (Invitrogen) according to the manufacturer’s protocols. A tobacco etch virus protease cleavage site was introduced into the N terminus of Cdc20. Sf9 insect cells were infected with the Cdc20 baculovirus and harvested at about 60 h post-infection. His6-tagged Cdc20–WD40 was purified with Ni2+-NTA agarose resin (Qiagen) and cleaved with tobacco etch virus protease to remove the His6-tag. The Cdc20–WD40 protein was further purified by anion exchange chromatography with a Mono-Q column followed by size exclusion chromatography with a Superdex 200 column (GE Healthcare). Purified Cdc20–WD40 was concentrated to 4–5 mg ml−1 in the Superdex 200 column buffer containing 25 mM Tris (pH 8.5), 150 mM NaCl, 1 mM MgCl2, 5% glycerol and 5 mM tris(2-carboxyethyl)phosphine.

The Cdc20–WD40 protein was crystallized at 20 °C using the sitting-drop vapour-diffusion method with a reservoir solution containing 0.1 M 2-(N-morpholino)ethanesulfonic acid (MES, pH 6.5), 15% (w/v) polyethylene glycol (PEG) 6000 and 5% MPD. The Cdc20–WD40 crystals were transferred to a new 2 µl sitting drop with a reservoir solution containing 0.1 M MES (pH 6.5) and 20% (w/v) PEG 6000 and soaked for 3–5 h to remove bound MPD. Apcin compound was dissolved into DMSO and added to the same drop at the final concentration of 5 mM. After overnight soaking, the crystals were cryo-protected in a solution containing 0.1 M MES (pH 6.5), 20% (w/v) PEG 6000, 10% glycerol and 5 mM apcin, then flash-cooled in liquid nitrogen. Crystals diffracted to a minimum Bragg spacing (dmin) of about 2.1 Å and exhibited the symmetry of space group P21 with cell dimensions of a = 41 Å, b = 87 Å, c = 48 Å and β = 113°, and contained one Cdc20 molecule per asymmetric unit.

Data collection and structure determination

Diffraction data were collected at beamline 19-ID (SBC-CAT) at the Advanced Photon Source (Argonne National Laboratory) and processed with HKL3000 (ref. 35). Phases were obtained by molecular replacement with Phaser using the crystal structure of human Cdc20–WD40 (Protein Data Bank accession number 4GGC) as search model36. Iterative model building and refinements used COOT and Phenix, respectively37,38. The final model for Cdc20–WD40–apcin (Rwork = 16.5%, Rfree = 21.3%) contained 313 residues, 82 water molecules and one apcin molecule. MolProbity was used for structure validation to show that all models had good geometry, except for one surface residue that was an outlier in a Ramachandran plot39. Data collection and structure refinement statistics are summarized in Extended Data Table 1.

High-throughput image-based assay to measure mitotic fraction

Parental A549, U2OS and hTERT-RPE1 cells were purchased from ATCC. DLD-1 cells were purchased from Sigma. For hTERT-RPE1, A549 and U20S cells, stable cell lines expressing H2B–GFP were derived using described methods40 and used in the experiments. DLD-1 cells were used without further modification. Cell lines were tested for mycoplasma contamination (Lonza kit LT07-218) after they were derived and were found negative. For each cell line, asynchronous cells were re-suspended to a density of 3.75 × 104 cells per millilitre. A WellMate dispenser (Thermo Scientific) was used to distribute 40 µl of suspension to each well of a black, clear-bottom 384-well plate (3712, Corning). Plates were sealed with breathable white rayon sealing tape (241205, Nunc) during plating and subsequent incubation. After 24 h incubation, the cells were treated with indicated concentrations of apcin and proTAME dissolved in DMSO, in four technical replicates. After 18 h, cells were fixed and stained directly without wash steps to avoid loss of mitotic cells, by adding 10 μl of 6× concentrated fixing/staining reagent (60% formalin, 0.6% Triton X-100 and 1.5 μg ml−1 Hoechst 33342 in DPBS). The plates were sealed with aluminium sealing tape (276014, Nunc) and incubated at 22–24 °C for 40 min before imaging. Plates were then imaged at four positions per well using an ImageXpress Micro (Molecular Devices) high-throughput microscope, with a ×10 objective, yielding a total of 16 images per condition (four images × four replicates). Cell images were processed automatically in ImageJ to identify the nuclei, count the number of nuclei and determine the maximum fluorescence intensity of each nucleus in each image. The output files from ImageJ for each treatment were pooled and the cumulative frequency curves of maximum intensity for the cell population in each treatment were computed using Matlab. An intensity threshold was set on the basis of the mitotic fraction in the wells treated with DMSO to separate mitotic cells from interphase cells. The interphase fraction for each treatment was indicated by the fraction below the threshold on the cumulative frequency plot. Methods for statistically analysing the data are presented in Supplementary Information.

Fluorescence live-cell imaging

For experiments with RPE1-H2B–GFP cells, asynchronous RPE1-H2B–GFP cells were plated in 24-well glass-bottom plates (Greiner BioOne, 662892) 18–24 h before siRNA transfection using RNAiMax (Invitrogen). Cells were transfected with Mad 2 siRNA (GGAACAACUGAAAGAUUGGdTdT, synthesized by Dharmacon) or non-targeting control siRNA (D-001210-01-20, Dharmacon) at a final concentration of 20 nM. Twenty-four hours after transfection, cells were treated with compounds and imaging was initiated following compound treatment. For U2OS-H2B–GFP cells, the cells were first synchronized by double thymidine block (18 h first block, 8 h release, 17 h second block; thymidine concentration 2 mM). Compounds were added at 7 h after release from the second block, and imaging was initiated after compound treatment. To measure efficiency of Mad2 knockdown by siRNA, western blot samples were each prepared from a single well of the 24-well glass-bottom plates. Twenty-four hours after transfection, cells were collected by trypsinization, pelleting and re-suspension in 2× NuPAGE sample buffer (Invitrogen) + 50 mM DTT.

For imaging, plates were inserted into a covered chamber supplied with humidified 5% CO2 and mounted onto a motorized microscope stage (Prior Scientific). Differential interference contrast and fluorescence images were captured at 6 min intervals for 45 h using a Nikon Ti inverted fluorescence microscope fitted with a 37 °C enclosed incubation chamber and using a ×20 Plan Apo 0.75 numerical aperture objective lens. A Hamamatsu ORCA cooled CCD (charge-coupled device) camera collected the images with 2 × 2 binning using Nikon Elements software (version 3.0). Videos were manually analysed using Nikon Elements software or ImageJ. Mitotic duration was defined as the time from nuclear envelope breakdown until anaphase, in the case of normal mitosis, or until exit from prolonged mitosis as indicated by cytoplasmic blebbing accompanied by changes in chromatin as detected by H2B–GFP. Cell fate was scored as ‘division’ if two daughter cells were produced by mitosis of any duration, and ‘abnormal exit’ if a single cell of interphase appearance resulted after that cell was in a mitotic state of any duration. The ‘death’ fate describes cells that entered mitosis and died while in mitosis. Methods for statistically analysing live cell imaging data are presented in Supplementary Information.