afm images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two heat-repeat...

34
AFM images of open and collapsed states of yeast condensin suggest a scrunching model for DNA loop extrusion Je-Kyung Ryu, 1 Allard J. Katan, 1 Eli O. van der Sluis, 1 Thomas Wisse, 1 Ralph de Groot, 1 Christian Haering, 2 and Cees Dekker 1, 3 1 Department of Bionanoscience, Kavli Institute of Nanoscience Delft, Delft University of Technology, Delft, Netherlands. 2 Cell Biology and Biophysics Unit, Structural and Computational Unit, European Molecular Biology Laboratory (EMBL), Heidelberg, Germany. 3 Corresponding author: [email protected] SUMMARY Structural Maintenance of Chromosome (SMC) protein complexes are the key organizers of the spatiotemporal structure of chromosomes. The condensin SMC complex, which compacts DNA during mitosis, was recently shown to be a molecular motor that extrudes large loops of DNA. The mechanism of this unique motor, which takes large steps along DNA at low ATP consumption, remains elusive however. Here, we use Atomic Force Microscopy (AFM) to visualize the structure of yeast condensin and condensin-DNA complexes. Condensin is found to exhibit mainly open ‘Oshapes and collapsed ‘Bshapes, and it cycles dynamically between these two states over time. Condensin binds double-stranded DNA via a HEAT subunit and, surprisingly, also via the hinge domain. On extruded DNA loops, we observe a single condensin complex at the loop stem, where the neck size of the DNA loop correlates with the width of the condensin complex. Our results suggest that condensin extrudes DNA by a fast cyclic switching of its conformation between O and B shapes, consistent with a scrunching model. KEY WORDS SMC protein, AFM microscopy, HS-AFM, DNA loop extrusion, condensin, scrunching model preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for this this version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358 doi: bioRxiv preprint

Upload: others

Post on 12-Sep-2020

1 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

AFM images of open and collapsed states of yeast condensin suggest

a scrunching model for DNA loop extrusion

Je-Kyung Ryu,1 Allard J. Katan,1 Eli O. van der Sluis,1 Thomas Wisse,1 Ralph de Groot,1

Christian Haering,2 and Cees Dekker1, 3

1Department of Bionanoscience, Kavli Institute of Nanoscience Delft, Delft University of

Technology, Delft, Netherlands.

2Cell Biology and Biophysics Unit, Structural and Computational Unit, European Molecular

Biology Laboratory (EMBL), Heidelberg, Germany.

3Corresponding author: [email protected]

SUMMARY

Structural Maintenance of Chromosome (SMC) protein complexes are the key organizers of the

spatiotemporal structure of chromosomes. The condensin SMC complex, which compacts DNA

during mitosis, was recently shown to be a molecular motor that extrudes large loops of DNA.

The mechanism of this unique motor, which takes large steps along DNA at low ATP

consumption, remains elusive however. Here, we use Atomic Force Microscopy (AFM) to

visualize the structure of yeast condensin and condensin-DNA complexes. Condensin is found

to exhibit mainly open ‘O’ shapes and collapsed ‘B’ shapes, and it cycles dynamically between

these two states over time. Condensin binds double-stranded DNA via a HEAT subunit and,

surprisingly, also via the hinge domain. On extruded DNA loops, we observe a single condensin

complex at the loop stem, where the neck size of the DNA loop correlates with the width of the

condensin complex. Our results suggest that condensin extrudes DNA by a fast cyclic switching

of its conformation between O and B shapes, consistent with a scrunching model.

KEY WORDS

SMC protein, AFM microscopy, HS-AFM, DNA loop extrusion, condensin, scrunching model

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 2: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

INTRODUCTION

The structural maintenance of chromosomes (SMC) family of protein complexes, such as

condensin, cohesin, and the SMC5/6 complex, is pivotal to cellular life (Aragon et al., 2013;

Hirano, 2016; Jeppsson et al., 2014; Nasmyth and Haering, 2005; Rowley and Corces, 2018;

van Ruiten et al., 2018; Uhlmann, 2016). These SMC protein complexes play vital roles for

spatially organizing chromosomes as they are key in sister chromatid cohesion, chromosome

condensation and segregation, DNA replication, DNA damage repair, and chromosome

resolution (Haering and Gruber, 2016; Nasmyth and Haering, 2005). Condensin plays a crucial

role to compact mitotic chromosomes during cell division (Hirano, 2016). Indeed, when

condensin is defective, both chromosome formation and the resolution of sister chromatids fail

(Houlard et al., 2015; Oliveira et al., 2005). Structurally, SMC protein complexes feature a

circularly connected conformation of a kleisin–SMC heterodimer ring (Figure 1A): Two anti-

parallelly folded coiled-coil arms, Smc2 and Smc4 (each ~45 nm in length), feature an

interconnected hinge domain at one end and heads with an ABC-type nucleotide-binding

domain at the other end, where the two head ends are mutually connected by a Brn1 kleisin.

Furthermore, two HEAT-repeat subunits, Ycg1 and Ycs4, are bound to the kleisin. A recent

crystal structure showed that DNA is anchored by Ycg1–Brn1 through a safety-belt mechanism

(Kschonsak et al., 2017).

Hi-C studies have suggested that mitotic chromosomes consist of DNA loops extruded by SMC

protein complexes (Dekker and Mirny, 2016; Fudenberg et al., 2016; Naumova et al., 2013;

Rao et al., 2017; Sanborn et al., 2015) In addition, polymer simulations indicated that loop

extrusion can result in the efficient disentanglement and compaction of chromatin fibers

(Alipour and Marko, 2012; Goloborodko et al., 2016). The formation of these loops, which can

be very long (up to Mbp), requires a processive motor that extrudes DNA. Recent single-

molecule fluorescence studies indeed demonstrated a motor function (Terakawa et al., 2017)

and provided direct evidence for DNA loop extrusion by the yeast condensin complex (Ganji

et al., 2018). Condensin constitutes a unique new type of DNA-translocase motor that reels in

DNA at high speed (up to 1.5 kbp/s at low forces) while consuming low ATP, i.e. an ATPase

rate of ~2 ATP/s, implying that the motor takes very large (~50 nm) steps per ATP molecule

that is hydrolyzed (Ganji et al., 2018). While this large step size suggests that the motor action

involves sizeable conformational changes at the level of the size of the full SMC complex, it is

currently unclear by what molecular mechanism condensin is able to process DNA so

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 3: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

efficiently. Various models have been suggested (Hassler et al., 2018). For example, in a

tethered inchworm model (Nichols and Corces, 2018), condensin moves along DNA by a

sequential binding of the heads to DNA, much like the motion of kinesin along microtubules.

In a DNA-pumping model (Diebold-Durand et al., 2017; Marko et al., 2019), a zippering of

two SMC coiled-coil arms pushes entrapped DNA from the hinge domain to the head domains,

while an ATP-induced dimerization of the head domains changes the condensin conformation

from an I-shape to O-shape to target new binding of DNA for subsequent cycles. In a scrunching

model (Terakawa et al., 2017), an extension and retraction of the flexible SMC arms move

DNA connected to the hinge domain to the globular subunits near the heads, and back in a

cyclic fashion. Although the open question on the motor mechanism is at the heart of resolving

the functioning of SMC proteins complexes, and hence is crucial for understanding

chromosomal structure, experimental evidence for any of these models has thus far been very

limited.

Studying the conformational changes of condensin and its connection to the mechanochemical

cycle that couples ATP hydrolysis to structural changes of the protein complex is thus critical

to the understanding of the mechanism underlying DNA loop extrusion by condensin. Many

studies have been undertaken to resolve the structure of a variety of SMC proteins, with

contradictory results so far, however (Eeftens and Dekker, 2017). Negative-staining EM

structures and crosslinking experiments suggested a rigid extended structure (I-shape) of

prokaryotic SMCs, while ATP-dependence studies indicated a transition from I- to O-shape

upon ATP hydrolysis (Diebold-Durand et al., 2017; Kamada et al., 2017; Soh et al., 2015). A

recent EM and cross-linking study on MukBEF and yeast cohesin suggested transitions between

an I shape and a folded state (Bürmann et al., 2019), similar to an early AFM study on S. pombe

condensin (Yoshimura et al., 2002). High-speed (HS) atomic force microscopy (AFM) images

of yeast condensin SMC dimers (without the kleisin and HEAT-repeat co-factors) in liquid

showed instead very flexible SMC arms and dynamical conformational changes between O-,

V-, P-, and B-shapes (Eeftens et al., 2016). However, thus far, no HS liquid AFM studies on

the full condensin holocomplex have been reported. High-resolution imaging of the yeast

condensin holocomplex with all its subunits is highly desirable in order to understand its

conformations at the stem of the DNA loop, where it reels in DNA.

Here, we resolve the conformational states and shape transitions of yeast condensin and its

binding configuration to DNA, with the aim to understand the motor dynamics that drive DNA

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 4: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

loop extrusion. We use both dry AFM and liquid HS AFM for visualization. AFM imaging

comes with significant advantages: it features a high resolution (xy resolution ~ 1 nm, z

resolution ~ 0.1 nm), it provides high contrast, and it is label free and cross-link free, making it

very suitable to study complex DNA-protein interactions. Dry AFM imaging, where the sample

is dried after incubating the molecules on a surface, enables high-throughput imaging by

capturing many molecules in various conformations, including intermediates of loop extrusion.

HS AFM in liquid even enables imaging of real-time structural changes of condensin within a

physiological buffer in the presence of ATP, with a temporal resolution up to 20 frames per

second, although it features a lower throughput due to small imaging area (~100×100 nm2) that

is needed to realize a high frame rate (Katan and Dekker, 2011; Kodera et al., 2010; Uchihashi

et al., 2011).

A major finding from our data is that the condensin holocomplex predominantly exhibits either

an open O-shape or a collapsed B-shape, and that it cycles back and forth between these states.

In addition, we find that the SMC arms are highly flexible and condensin binds to DNA through

both the hinge and HEAT-repeat subunits. This is suggestive of a scrunching model for this

DNA-loop-extruding motor, which involves a cycle where DNA is periodically transferred

between two different DNA-binding sites, near the hinge and near the globular HEAT/head

subunits.

RESULTS

Condensin exhibits an open O-shape or collapsed B-shape

Figure 1B shows an example of an AFM image of individual Saccharomyces cerevisiae

condensin holocomplexes, i.e., the full complex with all its 5 subunits. These purified proteins

showed DNA-stimulated ATPase activity and DNA loop extrusion (Figure S1), similar to

earlier reports (Ganji et al., 2018). The AFM images showed a homogeneous volume

distribution of the condensins (Figures 1B and S2A), with an average complex volume of 2,405

± 430 nm3 (N = 384). Similar to previous AFM studies that resolved the structure and various

conformational states of the arms of the ‘SMC dimer only’ Smc2–Smc4 complexes, we were

able to resolve the SMC arms in most of our condensin images, as well as distinguish various

conformational states of the condensin holocomplexes (Figures 1C–G).

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 5: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

The main configurations that condensin was found to adopt were O-shapes and B-shapes

(Figures 1C and D). In the O-shape, a ring-like configuration was clearly identified, with two

flexible arms, a large globular domain at one side, and a locally increased height at the hinge at

the other side. The globular domain is presumably composed of the two SMC head domains,

two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these

could not be resolved. In contrast, using liquid-phase AFM, we resolved multiple individual

subunits (Figure 1C): Smc2–Smc4 head domains and two HEAT-repeat subunits. The arms

were seen to adopt many different configurations, demonstrating a high flexibility. The O-

shapes exhibited on average a distance of 37 ± 7 nm (mean ± s.d.) between the center of mass

of the globular and hinge domains (Figure 1E), and featured a 102 ± 51 degrees opening angle

at the hinge (Figure 1F). The second prominent state of condensin that we observed was a

collapsed B-shape. This state appears to correspond to a ‘butterfly shape’ formed when the

hinge domain engages the joined SMC head domains, similar to observations in SMC-dimer-

only samples (Eeftens et al., 2016). To check whether surface charge affects the conformation

of the proteins, we also imaged complexes on polylysine-treated mica, which is a positively

charged surface, instead of on mica, which is slightly negatively charged (as in the data in

Figure 1). This yielded the same result (Figure S3), viz., a dominance of O shapes and B shapes,

indicating that these conformations are not induced by electrostatic surface-protein interactions.

Together, the O shapes and B shapes accounted for a high fraction (~80 %) of all data. Minority

states that were observed were I-shapes (9 %), kleisin-opened shapes (9 %), and hinge-opened

states (4 %). In the I-shapes, the two SMC arms are co-aligned, and the distance between hinge

and non-SMC subunits was slightly larger than the distance of O-shapes (Figure 1E). Finally,

we observed condensins with a disconnection between one of the SMC heads and the non-SMC

subunits (indicating a ‘kleisin-opened shape’; Figure 1C), and shapes where the two SMC arms

were disengaged at the hinge (‘hinge-opened shapes’). These, relatively rare, states where the

complex is partially opened may possibly relate to a topological loading of condensin onto

DNA, for which the ring has to open.

Different ATP states yielded a different distribution of O- and B-shapes. We performed AFM

experiments under several conditions: with ATP (N = 255), without ATP (N = 384), and using

an EQ mutant of condensin that binds ATP but is unable to hydrolyze it (N = 419). In the

presence of ATP (i.e., the data discussed above), there were slightly more B-shapes than O-

shapes. In contrast, significantly more open O-states were observed without ATP, while instead

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 6: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

more collapsed B-shapes were observed for the EQ mutant with ATP (Figure 1G). The latter

suggests that the collapsed B state may be promoted by ATP binding to the head domains,

whereas the observation that O states were more abundant without ATP may indicate that

hydrolysis and release of ATP are associated with a release of the hinge domain from the head

domains.

Condensin cycles back and forth between an open O-shape and collapsed B-shape

We applied HS AFM imaging to visualize the dynamics of the conformational changes of

condensin holocomplexes in a buffer with 50 mM NaCl and ATP – equivalent to the buffer

where condensin exhibits loop extrusion (Figure S1C; Ganji et al., 2018). Movies were recorded

at 5 frames per second (Figure 2; Movies S1–S4). We could resolve 3 or 4 substructures within

the globular domain, see e.g., Figure 2A and S4, where one can distinguish, next to the SMC

arms and hinge domain, the dimerized head domains and two HEAT-repeat subunits. The SMC

arms in the condensin holocomplex were found to be very dynamic and flexible, changing

conformations very fast (i.e., at a <0.2 s time scale in this example of a 5-Hz movie, Figure 2A),

very similar to previous studies on SMC-dimer-only samples (Eeftens et al., 2016). For the O-

shapes, we also observed that the hinge angle between the two SMC arms changed dynamically

(Figure S6), consistent with large conformational changes in the hinge domain.

Interestingly, we observed clear dynamic conformational changes where condensin would

toggle between an O-shape and a collapsed B-shape, and vice versa (Figures 2A–2B, and S4;

Movie S1-4). In these movies, the hinge region transitioned between a far-away position and a

location close to the globular domain, consistent with an overall shape change between the O-

and B-states, respectively (Figures 2A-2B). Quantification of this dynamics is provided in

Figure 2C, which plots the hinge-head distance as measured between their centers of masses

(cf. inset to Figure 2B). The data clearly exhibit a two-level-fluctuator behavior, displaying two

clear peaks in the histogram (right side of Figure 2B). The distance between these peaks,

averaged over multiple such traces (N = 12), was 14 ± 4 nm (mean ± s.d; Figure 2C). The

transition between the O-shape and the B-shape state can be completed very fast, faster than

our temporal resolution (0.2 seconds in this movie). Note that the hinge movement was not

strongly affected by tip-sample interaction, because, as shown in Figure S5, we find that the

hinge movement is isotropic and not biased by the AFM scanning direction, indicating that the

hinge displacement rather reflects the intrinsic behavior of this domain.

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 7: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Condensin binds double-stranded DNA via a HEAT subunit and via the hinge domain

In order to reel in DNA to extrude a DNA loop, the condensin motor complex should exhibit

multiple binding sites to DNA. Identification of the DNA binding sites is therefore important.

We used dry AFM to examine the interaction between condensin and DNA. We imaged large

scan areas (>10×10 μm2) in order to gather good statistics in counting bound condensin

complexes. The AFM images showed different binding modes (Figure 3). First, as expected,

we observed that condensin binds to DNA at the globular domain (Figure 3A), presumably at

the Ycg1–Brn1 interface (Kschonsak et al., 2017). Second, we observed binding of condensin

to DNA via the hinge domain (Figure 3B). Unexpectedly, hinge binding was observed to occur

very frequently, almost as often as HEAT binding, viz., for 35 versus 33 % of the complexes,

respectively (Figure 3E). Third, we observed that condensin can bind DNA through both the

hinge and globular domains (Figure 3C). Condensin was also observed to bind to the DNA in

the collapsed B-state (26%; Figure 3D), but here it was difficult to discriminate the exact

binding location within the complex.

A number of controls supported these results. Firstly, we established that an accidental

colocalization, i.e. a possible fortuitous overlap of DNA and unbound condensin, contributes

only minor fraction in the presence of ATP (Figure S7), indicating that condensin binding to

DNA is driven by biochemical interactions. Secondly, the DNA-binding probability of wild-

type condensin was much higher than that for a complex that lacks the Ycg1 subunit (Figure

S7B), consistent with a previous study that Ycg1 is important for DNA binding (Piazza et al.,

2014). Finally, we observed a decrease in the number of condensins binding to DNA in

experiments where we omitted ATP or replaced it with ATPγS (Figure S7B), consistent with

previous observations that ATP hydrolysis is needed to topologically load condensin onto DNA

(Eeftens et al., 2017) and increase condensin’s residence time on DNA (Ganji et al., 2018).

DNA loops feature a single condensin complex at their stem

Next, we used AFM to image condensin holocomplexes at the stem of loops of DNA that were

formed by condensin. To observe such extruded loops in dry AFM images, it is important to

avoid stretching forces on the DNA during the drying process, since forces as low as ~10 pN

are enough to disrupt the loops formed by condensin (Eeftens et al., 2016). Using a careful

drying procedure (see Materials and methods), we could image condensins on extruded loops

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 8: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

of DNA with dry AFM. To prepare samples, we mixed condensin with DNA and ATP in a test

tube and incubated the mixture for 3 min, an incubation time similar to the lag time between

condensin arrival and the initiation of DNA compaction observed in a magnetic tweezers

experiment (Eeftens et al., 2017). For dry AFM imaging, we subsequently deposited the

samples under a very slight stretching force onto a mica surface that was pre-functionalized

with a low concentration of polylysine, while for liquid-phase AFM, we immobilized the

sample onto mica with the addition of 0.3 mM spermidine.

Both in liquid and dry AFM, we observed a clear co-localization of condensin and the stem of

DNA loops (Figures 4A-B; Movie S5). We conclude that these loops were predominantly

formed by condensin in an ATP-hydrolysis-dependent manner, based on the following

observations. First, the number of co-localizations was much higher than could be expected

from random colocalization (Materials and Methods; Figures S8). Second, the loop size of DNA

was found to be significantly larger for the case of incubation with ATP than for any negative

controls (Figures 4E and S9). Moreover, the number of condensin-mediated loops depended

strongly on the availability of hydrolysable ATP (Figure 4C). Negative controls with ATPγS or

without nucleotide showed a significantly lower number of loops per unit DNA length, 0.029

loops/μm and 0.019 loops/μm respectively, compared to 0.15 loops/μm for the case with ATP

(counting, in this case, only loops that have a condensin at their stem; see also Figure S8D).

And finally, we rarely observed loops when using a tetrameric a S. cerevisiae condensin

complex that lacks Ycg1 (Figure 4C). Taken together, these experiments show that condensin

localizes to the loop stem when extruding DNA loops.

We quantified the number of condensin molecules at the stem of the DNA loops by estimating

their volume. We found that 90 % of the DNA loops showed a single condensin at the loop

stem (Figure 4E), consistent with an earlier estimate from optical measurements (Ganji et al.,

2018) and confirming that a single condensin holocomplex is responsible for loop extrusion of

the DNA. We rarely (10%) observed multimeric forms of condensin at the stem of the DNA

loops in our experimental conditions. Our observation of single condensins at the DNA loop

stem contrasts suggestions of the handcuff model where loop extrusion by cohesin or MukBEF

was suggested to be driven by dimeric SMC complexes (Woo et al., 2009; Zhang et al., 2008).

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 9: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

The neck width of the DNA loop stem correlates to the size of the condensin complex

Next, we examined the conformation of protein and DNA at the loop stem in more detail.

Specifically, we examined the correlation between the neck width at the stem of the DNA loop

(Figure 5) and the width of the condensin molecule located there. The neck width of the DNA

loop was determined from the distance between the two pieces of DNA right next to the

condensin at the loop stem, as measured from the height profile of the cross section across the

two pieces (Figure 5A, bottom). The condensin width was similarly obtained from the height

profile of the cross section. For good comparison, both the DNA neck width and protein width

were defined as the distances between two outer points of the full width half maximums of the

Gaussian fits, since separated peaks could not always be resolved when the two DNA molecules

were too close.

Interestingly, we found a clear linear relationship between the width of the DNA loop at the

stem and the characteristic size of the protein, as measured at the neck (correlation coefficient

R2 = 0.87; Figures 5B and S10C). This strong tendency to follow the protein conformation

suggests that DNA is bound to a condensin complex that varies in size. Next to the population

that showed this linear correlation, a second smaller population (blue data in Figure 5B; and

S10D) showed no dependence of the DNA neck width on condensin size. From studying the

individual images (Figure S10), we conclude that the small DNA-neck width of these molecules

is caused by interwound structures within the DNA, that likely are induced upon depositing the

DNA onto the polylysine-treated mica surface (Japaridze et al., 2017a). Overall, the data

suggest that conformational changes of the condensin complex appear to be associated with

changes in the width of the stem of the DNA loop that is extruded by condensin.

DISCUSSION

Summing up, we visualized yeast condensin holocomplexes on their own as well as bound to

DNA. We found that condensin predominantly exhibits an O-shape or collapsed B-shape, and

that it dynamically switches between these conformations. When condensin was added to DNA,

it bound at the hinge or at the globular domain that is formed by the SMC heads and HEAT-

repeat subunits. Condensin formed DNA loops, and we observed that a single condensin was

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 10: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

located at the loop stem, where the loop-stem width correlated with the condensin size. Below

we discuss the relevance and importance of these findings.

Flexible SMC arms

Our liquid HS-AFM imaging showed very flexible SMC arms with highly dynamic

conformational changes. This is very similar to previous observations for coiled coil arms of

the SMC dimers only, which exhibited a persistence length of only 4 nm (Eeftens et al., 2016),

indicating that the flexibility of SMC arms is a general feature of the condensin holocomplex

and not influenced by the interaction between SMC arms and non-SMC subunits. While the

HS-AFM images unambiguously show flexible SMC arms, more rigid SMC arms were implied

in reports using other techniques, such as X-ray crystallography, cross-linking experiments, or

EM (Diebold-Durand et al., 2017; Soh et al., 2015). For Rad50, a related SMC-like protein,

liquid phase imaging similarly showed flexible and dynamical SMC arms (Anderson et al.,

2001; Moreno-Herrero et al., 2005), while crystal structures showed rigid rods (Park et al.,

2017). HS-AFM thus appears to be a liquid-phase imaging technique that can visualize

dynamical changes of the SMC complexes with high spatial and temporal resolution, which

may be missed in X-ray crystallography, cross-linking experiments, and EM imaging of fixed

samples.

Conformation of the condensin complex

Our data show that condensin predominantly exhibits an extended O-shape or collapsed B-

shape. This fits observations in an early AFM study and a recent EM paper that showed both

extended and compacted configurations of SMC proteins (Yoshimura et al., 2002; Bürmann et

al., 2019). However, our data did not show a juxtaposed I-shape that was reported in these and

other studies (Soh et al., 2015; Diebold-Durand et al., 2017). There are several possible

explanations for the discrepancy of O- versus I-shape observations among various studies. First,

SMC complexes from different organisms may exhibit different shapes. For example, S.

cerevisiae cohesin and non-crosslinked MukBEF were found to adopt a folded conformation

(Bürmann et al., 2019), while B. subtilis SMC proteins showed I-shapes and O-shapes (Diebold-

Durand et al., 2017). Second, the condensin complex is a large protein complex and purifying

intact functional holocomplexes is challenging. We verified that our proteins were functional

and capable of loop extrusion (Figure 4 and S1). Third, different sample-preparation methods

can induce or bias the observation of certain configurations. The high signal-to-noise ratio in

AFM imaging enables to analyze every single molecule in the field of view, which prevents the

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 11: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

problem of selection bias that is an ever-luring danger in single-particle electron microscopy

(Lyumkis, 2019). Furthermore, the air-water interface in cryo-EM samples can cause a

preferred orientation of proteins (Noble et al., 2018) or denature proteins (D’Imprima et al.,

2019). The sample preparation for AFM is relatively mild and no staining or labeling is

required. From the AFM imaging in solution, it is clear that the coiled-coil arms and hinge

domain are parts of the complex that interact only weakly with the surface, which facilitates the

condensin to adopt an equilibrated shape. However, the presence of a surface is a requirement

in AFM (as also is the case in most EM techniques), which potentially may affect the condensin

structure. We measured both on slightly negatively charged mica surfaces and on positively

charged polylysine-treated surfaces (Figure S3), and O-shapes and B-shapes were the most

abundant in both cases, suggesting that these shapes are intrinsic to condensin.

DNA binding sites of the condensin complex

Our AFM images of DNA-bound condensin show that condensin can bind DNA at (at least)

two binding sites, one at the hinge and one at the globular domain with the SMC heads and

HEAT-repeat subunits. Previous work clearly showed that Ycg1-Brn1 provides a strong DNA

binding site (Kschonsak et al., 2017), which is consistent with the binding that we observed

near the globular domain that includes the HEAT-repeat subunits (Figure 3A). The frequent

occurrence of hinge binding that we observed is more surprising. Since the inner region of the

hinge domain has a positively charged surface, it is actually plausible that this may bind DNA

(Griese et al., 2010). Indeed, gel-shift assays and fluorescence anisotropy experiments showed

DNA binding affinity to the hinge domain of mouse condensin (Griese et al., 2010), yeast

condensin (Piazza et al., 2014), bovine cohesin (Chiu et al., 2004) and prokariotic condensin

(Hirano, 2016). Previous AFM studies also showed DNA binding by the hinge domain for

MukB (Kumar et al., 2017) and for S. pombe condensin (Yoshimura et al., 2002), suggesting a

conserved hinge-binding mechanism. Our observation that the hinge angle of the O-shaped

condensin changes dynamically (Figure S6) indicates that the hinge domain is very flexible.

This is consistent with previous crystal structures of hinge domains that showed two different

states of the fully associated and half-dissociated conformations (Griese et al., 2010; Haering

et al., 2002; Ku et al., 2010; Li et al., 2010; Niki, 2017).

Implication for models of DNA loop extrusion by SMC complexes

Given that it thus far has remained puzzling how the condensin complex is able to extrude a

DNA loop, it is of interest to discuss what our findings imply for a mechanistic model for SMC-

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 12: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

induced DNA loop extrusion and to see how it relates to various models that have been

proposed. A working model should account for a number of specific experimental observations:

it should (i) involve cyclic conformational transitions between an O-shape and B-shape of the

SMC complex, (ii) be compatible with very flexible SMC arms, (iii) accommodate a very large

step size (Ganji et al., 2018), (iv) show a high correlation between the condensin size and the

neck width at the stem of DNA loops (Figure 5), (v) use Ycg1-Brn1 as an DNA anchoring site

(Ganji et al., 2018; Kschonsak et al., 2017), and (vi) use a DNA-binding site at the hinge

domain.

Our experimental results appear to rule out the ‘tethered inchworm model’ (Nichols and Corces,

2018), which predicts that condensin moves along DNA by a sequential binding of the heads,

similar to the motion of kinesin along microtubules. Instead of the prominent head-head

motions involved in this model, we observed substantial hinge-head motions. Furthermore, this

model assumes that the hinge domain anchors the DNA, which is contradictory to the finding

that Ycg1-Brn1 is a DNA anchoring site. Furthermore, the ‘DNA pumping model’(Diebold-

Durand et al., 2017) is also inconsistent with our observations. In the DNA pumping model, a

zipping of the SMC arms from an O- to I-shape pushes DNA from the hinge to the head

domains, whereupon a subsequent conformational change from I- to O-shape targets new DNA.

The I-shapes thus play a prominent role in this model, whereas a B-shape does not occur in this

cycle – both in stark contrast to our observations.

Our data instead indicate that some form of a ‘scrunching model’ underlies the motor action of

SMCs (Figure 6), where condensin anchors to DNA and moves another piece of the DNA

relative to this point to extrude a loop, by cyclically binding DNA to the hinge and reeling it to

the globular heads/non-SMC domain. In such a model, the hinge binds DNA, transfers it to the

head region, binds it there, upon which the hinge releases to subsequently grab a new piece of

DNA to reel it in (Terakawa et al., 2017; Hassler et al., 2018). Our observation that condensin

dynamically switches between an O-shape or collapsed B-shape clearly hints at such a type of

mechanism. Our finding that the hinge binds DNA further supports this model. The flexible

SMC arms provide a hint on the mechanism of DNA transfer from the hinge to the globular

domain: While such flexible arms prevent a rigid-body power-stroke motion to directly

transduce mechanical motion from the ATPase head domain to the distant (37 nm) hinge

domain, the arms instead may use thermal fluctuations to facilitate the motion of the DNA-

bound hinge region to the heads/non-SMC subunits as a biased-ratchet motor (Hwang and

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 13: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Karplus, 2019). ATP binding may induce the interaction between the hinge and the head

domains to form a B-shape, whereas subsequent ATP hydrolysis may release the hinge (but not

the DNA) from the globular domain to restart the cycle. During this cycle, the Ycg1–Brn1

anchor would ensure that condensin remains locked to DNA, such that the newly reeled-in DNA

indeed translocates with respect to the anchor position, thus forming a DNA loop.

Such a scenario does imply the presence of another DNA binding site (next to the hinge and the

Ycg1–Brn1 anchoring site) in the globular heads/non-SMC subunits that prevents slipping of

the extruded DNA loop while the hinge is released in search of another piece of DNA. Based

on previous suggestions for Rad50 and prokaryotic SMC complexes (Erickson, 2009; Liu et

al., 2016; Seifert et al., 2016; Woo et al., 2009), such a third DNA-binding site might be located

in a positively charged cavity of the ATP-dimerized head domains or in the head-proximal

coiled coils (Rojowska et al., 2014). Furthermore, it is still an open question how the

unidirectionality of the loop extrusion is established in the scrunching model. This may relate

to the angle with which condensin binds DNA at its (transient) binding site, but resolving this

will have to wait for further structural studies.

In conclusion, our experimental observations support a scrunching model in which a DNA-

bound condensin extrudes a loop of DNA by cycling its conformation between O- and B-

shapes. It will be of interest to explore whether such a motor mechanism also applies to other

SMC complexes such as cohesin (Davidson et al., 2019; Kim et al., 2019).

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 14: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

ACKNOWLEDGMENTS

We thank S. Bisht, J. Eeftens, M. Ganji, E. Kim, B. Pradhan, I. Shaltiel, J. van der Torre, and

W.W.W. Yang for discussions. This work was supported by ERC grants SynDiv 669598 (to

C.D.), Marie Sklodowska-Curie grant agreement No 753002. (to J.-K.R.), CondStruct 681365

(to C.H.H.), the Netherlands Organization for Scientific Research (NWO/OCW) as part of the

Frontiers of Nanoscience and Basyc programs, and the European Molecular Biology Laboratory

(EMBL).

AUTHOR CONTRIBUTIONS

J.-K.R., C.D., A.J.K., and C.H.H. designed the experiments, J.-K.R., A.K., and T.W., performed

the AFM experiments, E.V.D.S. and J.-K.R. purified condensin complex, J.-K.R., A.J.K., and

T.W. contributed image analysis, J.-K.R and R.D.G. performed ATPase assay and single-

molecule fluorescence assay, C.D. and C.H.H supervised the work, J.-K.R., C.D., and A.J.K.

wrote the manuscript.

COMPETING INTERESTS

The authors declare no competing interests

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 15: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

MATERIALS and METHODS

Condensin holocomplex purification

We used the same protocol as previously reported (Ganji et al., 2018) for the purification of S.

cerevisiae condensin holocomplexes with all the subunits: Wild-type (Smc2-Smc4-Brn1-Ycs4-

Ycg1), Ycg1-missing mutant (Smc2-Smc4-Brn1-Ycs4), and EQ mutant (Smc2E1113Q-

Smc4E1352Q-Brn1-Ycs4). The transformation of these complexes was done using using 2μ-based

high copy plasmids containing pGAL10-YCS4 pGAL1-YCG1 TRP1 (Ycg1-missing mutant) and

pGAL7-SMC4-StrepII3 pGAL10-SMC2 pGAL1-BRN1-His12-HA3 URA3 (wild-type, strain

C4491). The overexpressed cells were lysed in buffer A (50 mM TRIS-HCl pH 7.5, 200 mM

NaCl, 5% (v/v) glycerol, 5 mM β-mercaptoethanol, 20 mM imidazole) supplemented with 1×

complete EDTA-free protease inhibitor mix (11873580001, Roche) in a FreezerMill (Spex),

cleared by centrifugation, loaded onto a 5-mL HisTrap column (GE Healthcare) and finally

eluted with 220 mM imidazole in buffer A. Eluate fractions were supplemented with 1 mM

EDTA, 0.2 mM PMSF and 0.01% Tween-20, incubated overnight with Strep-Tactin Superflow

high capacity resin (2-1208-010, IBA), and eluted with buffer B (50 mM TRIS-HCl pH 7.5,

200 mM NaCl, 5% (v/v) glycerol, 1 mM DTT) containing 10 mM desthiobiotin. After

concentrating the eluate by ultrafiltration, final purification proceeded by size-exclusion

chromatography with a Superose 6 column (GE Healthcare) pre-equilibrated in buffer B

containing 1 mM MgCl2. Purified protein was snap-frozen and stored at –80°C until use. Most

of the time, however, we used fresh condensin that had not been frozen.

ATPase assay

We used a colorimetric phosphate detection assay (Innova Biosciences) to measure the ATPase

activity of condensin complex. We mixed 50 nM condensin complexs and 50 ng/uL labmda

DNA (Promega) for 15 mins in 40 mM Tris-HCl (pH7.5), 50 mM NaCl, 2.5 mM MgCl2, 2 mM

DTT, 5 mM ATP. After 15 min incubation at room temperature, the released phosphate’s

concentration was measured following the protocol of the manufacturer.

Sample preparation for dry AFM

For the condensin-only sample preparation, we deposited 5 nM condensin onto newly cleaved

mica surface (or 0.00001 % polylysine pretreated mica in Figure S3) for various conditions,

namely with/without ATP and for EQ mutant condensin holocomplex with ATP in B1 buffer

(20 mM Tris, 50 mM NaCl, 2.5 mM DTT, and 2.5 mM MgCl2). We used a 2.5 mM ATP

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 16: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

concentration for WT/ATP or EQ/ATP conditions, and incubated the mixtures for 10 min. After

depositing the sample for 20 seconds, we rinsed the mica using H2O, and dried the sample by

using a gentle stream of N2. Then, we classified the structures into O-, I-, B-shapes, or kleisin-

or hinge-opened states.

For the study on DNA and condensin interactions, we mixed 3 ng/μL of lambda DNA

(D1501, Promega) and 5 nM of condensin in an Eppendorf tube for a 10 min incubation to

induce condensin-DNA interaction. We then added 1 mM ATP and incubated the sample for

an additional minute. We deposited the samples onto 0.00001 % (w/v) polylysine-treated mica

for 20 seconds and rinsed the surface using 3 mL MilliQ water. Finally, we dried the sample

using N2 gas.

To induce DNA loop formation by condensin, we incubated 3 ng/μL of lambda DNA

with 1.5 nM condensin for 5 min in B1 buffer, and added 100 μM of ATP to this solution. We

then incubated the mixture for 4 min. We deposited the sample (volume of 7 μL) onto the

polylysine-treated mica for 20 sec, and rinsed the sample using 1 mL of buffer containing 50

mM MgCl2 to completely immobilize the DNA on the surface and prevent DNA diffusion

during the washing and drying step. Then, the sample was rinsed with 1 to 3 mL MilliQ water.

While rinsing the sample on the mica, we tilted the mica with a small angle (~10° with respect

to horizontal direction) in order to reduce tension on the DNA sample. The sample was dried

by a gentle stream of N2.

Upon using high polylysine concentrations, DNA is strongly attached on the surface as

soon as it deposited, which yields kinetic trapping (Rivetti et al., 1996). The effective

persistence length for such a projection is smaller than for an equilibrated structure that is

obtained for DNA on MgCl2-treated mica. To prevent a high density of kinetically trapped DNA

loops, we minimized the polylysine concentration to 0.00001 % (w/v), so that DNA was fixed

onto the surface only intermittently using sparsely distributed polylysine molecules.

Dry AFM imaging

We performed our AFM measurements in air on a Bruker Multimode AFM, with a Nanoscope

V controller and Nanoscope version 9.2 software. We used Bruker ScanAsyst-Air-HR

cantilevers (nominal stiffness and tip radius 0.4 N/m,and 2 nm, respectively) or Bruker

Peakforce-HIRS-F-A ( 0.35 N/m and 1 nm). The imaging mode we used was PeakForce

Tapping, with an 8 kHz oscillation frequency, and a peak force setpoint value less than 100 pN.

For imaging the condensin structures, 3 μm × 3 μm scan areas with 2,048 × 2,048 pixels were

recorded at 0.5 Hz scanning speed. To record condensin on DNA loops, images of 10 μm × 10

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 17: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

μm were imaged with 5,120 × 5,120 pixels at 0.2 to 0.7 Hz scanning speed. All measurements

were performed at room temperature.

Liquid AFM imaging

Freshly purified condensin holocomplex (2 nM) that had not been frozen was deposited on

freshly cleaved mica surface with an imaging buffer (20 mM Tris [pH7.5], 50 mM NaCl, 1 mM

ATP, 2.5 mM MgCl2, 2.5 mM DTT). After 10 s, the sample surface was rinsed with the imaging

buffer. During the rinsing step, the sample was not dried. Following a published protocol

(Uchihashi et al., 2012), we imaged the sample with HS-AFM (HS-AFM 1.0, RIB) using

Nanoworld SD-S-USC-f1.2-k0.15 (2 nm tip radius, k = 0.15 N/m, f = 1.2 MHz) or Nanoworld

SD-S-USC-f1.5-k0.6 cantilevers (2 nm tip radius, k = 0.6 N/m, f = 1.5 MHz). Typically, a scan

size of 100 nm × 100 nm and 150 scan lines were used, with 2-10 Hz frame rates. For

minimizing and stabilizing the sample-tip interaction during imaging, we used a feedback

mechanism on the second harmonic amplitude (Schiener et al., 2004). We reconstructed movie

files and images using a Matlab script. Using a Matlab code, we measured the hinge-head

distance of every frame by measuring the distance between the two centers of mass of hinge

and head domains. The hinge angle was measured as the angle between the two SMC arms at

the hinge domain.

Image processing

Before quantitative analysis, images were processed to remove background and transient noise

data that would give false signals in automated analysis. This was done using Gwyddion version

2.53 (Nečas and Klapetek, 2012). To ensure that only the empty surface was used for

background subtraction, an iterative procedure of masking particles and subtracting (planar

and/or line-by-line) background polynomials was employed. Horizontal scars, which occur

from time to time due to feedback instabilities or particles sticking to the AFM tip, were selected

and removed by Laplacian background substitution. Finally, we used the blind tip estimation

and surface reconstruction algorithms in Gwyddion to reduce the effects of AFM tip

convolution (the widening of features due to the finite size of the AFM tip) on our images. It

has been shown that the tip shape can affect estimates of SMC protein volume (Fuentes-Perez

et al., 2012). Although blind tip estimation is not capable of fully correcting for the influence

of tip shape, we found that it significantly improved the consistency of the volume estimates,

and the increased contrast in the reconstructed images allowed for a better delineation of

individual molecules in automated image quantification.

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 18: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Characterizing the condensin complex from dry AFM images

The volume of condensin molecules was measured using a home-built Matlab code.

Background noise was subtracted by the program, and both the area and mean height of the

structure was measured with respect to the background surface. The volume was then calculated

by multiplying the deduced area and height. The distributions of the volumes were confirmed

by Gywddion. To measure the volume using Gywddion, a height mask was applied to cover the

condensin and grain measurement was used in order to determine the volume for the different

masked objects.

Distances between the hinge-nonSMC subunits were measured between the hinge and

nonSMC/head globular parts, using the line tool and measurement function of image J. The

hinge angle was determined by measuring the angle between two SMC arms at the hinge

domain.

Measurement of DNA length and DNA loop size

We wrote Matlab scripts that perform automated image analysis to deduce the path of the DNA

in images. The following procedure was used to select pixels that belong to areas covered with

DNA: The image was smoothed with a Gaussian filter (𝜎 = 0.5 pixels, window size = 5 pixels),

and pixels with a height of more than 0.15 nm were selected. To get rid of noise and

contaminations, objects (contiguous selected areas) with a maximum height less than 0.3 nm

were removed from the selection, as well as objects of area less than 1000 nm2. To selectively

analyze DNA on the surface, we excluded DNA-unbound proteins on the background by

excluding objects with boundary length-to-area ratio > 0.1 nm-1 and those with area-to-

bounding-box-area ratio < 0.25. As there was virtually no difference in observed height between

the SMC arms and the DNA, condensin molecules touching the DNA were mostly included in

the selection, and the total DNA length estimate is therefore a slight overestimate. The objects

selected in this way were identified to be DNA, and visual inspection confirmed that this was

reliable. To subsequently obtain the DNA length, the objects were skeletonized and the length

was approximated as 𝑑𝑝𝑛𝑠, with 𝑛𝑠 the number of pixels of the skeleton, and 𝑑𝑝 the linear size

of a single image pixel. For measuring the size of individual loops, loops were identified

manually and subsequently traced using DNA Trace software (Japaridze et al., 2017b).

Estimation of the probability of accidental co-localization of condensin and DNA

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 19: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

The binding probability between condensin and DNA was estimated from AFM images as the

fraction of proteins that overlapped with a DNA molecule. There is of course also a probability

that a randomly deposited protein will overlap with DNA. This accidental co-localization

probability is estimated as follows: We approximate the protein by its circumcircle with

diameter Dp. To overlap with the DNA, the center of this circumcircle should be within a

distance Dp/2 from the DNA. With a random placement of a condensin complex to the surface,

the probability of this occurring in co-localization with the DNA is equal to the relative area of

the image occupied by the space between two curves parallel to the DNA with a distance Dp

between them. If the total length of all DNA in the image is LDNA, this area can be approximated

by Dp×LDNA. See Figure S7A for a graphical illustration of this. Therefore, the probability of

finding a randomly placed molecule co-localized with the DNA is

𝑃𝐶,𝑟𝑎𝑛𝑑𝑜𝑚 = 𝐿𝐷𝑁𝐴×𝐷𝑝

𝐴 𝑖 ,

where 𝐴𝑖 is the imaged area. We used a value of 40 nm for Dp, and with an average LDNA = 209

μm DNA in a 10×10 μm2 image, the accidental colocalization probability is about 8%. Note

that this is an overestimate, since the condensin complex is often elongated so that in many

cases the circumcircle can touch the DNA but the protein does not, and furthermore the DNA

is curved and as a result the area between the parallel curves is less than Dp×LDNA.

A similar line of reasoning can be applied to the localization of condensin at the loop

stem (Figures S8A-S8C). Long DNA molecules that are attached to a surface have loops in

them even in the absence of protein, simply due to the deposition of the randomly coiled DNA

to the surface. There is a small chance that a randomly deposited protein, that was not already

bound to the DNA before, locates with a distance Dp from the loop crossing – a probability that

can be calculated similar to above. Furthermore, a condensin that was already bound to the

DNA before it got deposited, can, by chance, get located at the loop crossing in the process of

depositing the DNA to the surface. Combining these two terms leads to an expression for the

probability that condensin is accidentally observed as co-localized with loop crossings, which

equals the number of DNA-bound condensins that are randomly located at the stem of DNA

loops divided by the total number of DNA-bound condensins,

𝑃𝑙,𝑟𝑎𝑛𝑑𝑜𝑚 =2𝑁𝑙𝐷𝑝

𝐿𝐷𝑁𝐴+

𝜋𝐷𝑝2𝑁𝑙𝑁𝑝

4𝐴𝑖𝑁𝑏 ,

where 𝑁𝑙, 𝑁𝑝 and 𝑁𝑏 are the total number of loops, total number of proteins, and the number of

DNA-bound proteins in the image, respectively. For our typical conditions, this probability

𝑃𝑙,𝑟𝑎𝑛𝑑𝑜𝑚 is estimated to be about 3.5 %. See also Figure S8B.

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 20: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Width distributions of DNA-loop neck and protein

The neck size of the DNA was measured using the height profile tool of image J. We used the

cross-section of the DNA region closest to the loop neck that did not overlap with the site of

the protein. The DNA height profiles of the cross-section region showed two peaks, one from

each side of the loop forming the neck, to each of which we fitted a Gaussian. The neck size

was taken as the horizontal distance between the outer points of the FWHM of these Gaussians.

The distance was defined this way as it gives a consistent definition, both when the DNA

molecules or parts of the condensin complex were or were not resolved as separate Gaussian

peaks. The same procedure was followed for measurements of the width of the protein that was

located at the neck site.

Loop extrusion visualization using single-molecule fluorescence microscopy

For the loop extrusion assay, we followed the protocol published by (Ganji et al., 2018).

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 21: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

REFERENCE

Alipour, E., and Marko, J.F. (2012). Self-organization of domain structures by DNA-loop-

extruding enzymes. Nucleic Acids Research 40, 11202–11212.

Anderson, D.E., Trujillo, K.M., Sung, P., and Erickson, H.P. (2001). Structure of the

Rad50·Mre11 DNA Repair Complex from Saccharomyces cerevisiae by Electron

Microscopy. Journal of Biological Chemistry 276, 37027–37033.

Aragon, L., Martinez-perez, E., and Merkenschlager, M. (2013). Condensin, cohesin and the

control of chromatin states. Current Opinion in Genetics and Development 23, 204–211.

Bürmann, F., Lee, B., Than, T., Sinn, L., O’Reilly, F.J., Yatskevich, S., Rappsilber, J., Hu, B.,

Nasmyth, K., and Löwe, J. (2019). A folded conformation of MukBEF and cohesin. Nature

Structural & Molecular Biology 26, 227–236.

Chiu, A., Revenkova, E., and Jessberger, R. (2004). DNA interaction and dimerization of

eukaryotic SMC hinge domains. Journal of Biological Chemistry 279, 26233–26242.

D’Imprima, E., Floris, D., Joppe, M., Sánchez, R., Grininger, M., and Kühlbrandt, W. (2019).

Protein denaturation at the air-water interface and how to prevent it. ELife 8, 1–18.

Davidson, I.F., Bauer, B., Goetz, D., Tang, W., Wutz, G., and Peters, J. (2019). DNA loop

extrusion by human cohesin. Science.

Dekker, J., and Mirny, L. (2016). The 3D Genome as Moderator of Chromosomal

Communication. Cell 164, 1110–1121.

Diebold-Durand, M.L., Lee, H., Ruiz Avila, L.B., Noh, H., Shin, H.C., Im, H., Bock, F.P.,

Bürmann, F., Durand, A., Basfeld, A., et al. (2017). Structure of Full-Length SMC and

Rearrangements Required for Chromosome Organization. Molecular Cell 67, 334-347.e5.

Eeftens, J., and Dekker, C. (2017). Catching DNA with hoops—biophysical approaches to

clarify the mechanism of SMC proteins. Nature Structural & Molecular Biology 24, 1012–

1020.

Eeftens, J.M., Katan, A.J., Kschonsak, M., Hassler, M., de Wilde, L., Dief, E.M., Haering,

C.H., and Dekker, C. (2016). Condensin Smc2-Smc4 Dimers Are Flexible and Dynamic. Cell

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 22: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Reports 14, 1813–1818.

Eeftens, J.M., Bisht, S., Kerssemakers, J., Kschonsak, M., Haering, C.H., and Dekker, C.

(2017). Real‐ time detection of condensin‐ driven DNA compaction reveals a multistep

binding mechanism. The EMBO Journal 36, 3448–3457.

Erickson, H.P. (2009). Size and shape of protein molecules at the nanometer level determined

by sedimentation, gel filtration, and electron microscopy. Biological Procedures Online 11,

32–51.

Fudenberg, G., Imakaev, M., Lu, C., Goloborodko, A., Abdennur, N., and Mirny, L.A.

(2016). Formation of Chromosomal Domains by Loop Extrusion. Cell Reports 15, 2038–

2049.

Fuentes-Perez, M.E., Gwynn, E.J., Dillingham, M.S., Moreno-Herrero, F., Fuentes-Perez,

M.E., Gwynn, E.J., Dillingham, M.S., and Moreno-Herrero, F. (2012). Using DNA as a

fiducial marker to study SMC complex interactions with the atomic force microscope.

Biophysical Journal 102, 839–848.

Ganji, M., Shaltiel, I.A., Bisht, S., Kim, E., Kalichava, A., Haering, C.H., and Dekker, C.

(2018). Real-time imaging of DNA loop extrusion by condensin. Science 360, 102–105.

Goloborodko, A., Marko, J.F., and Mirny, L.A. (2016). Chromosome Compaction by Active

Loop Extrusion. Biophysical Journal 110, 2162–2168.

Griese, J.J., Witte, G., and Hopfner, K.P. (2010). Structure and DNA binding activity of the

mouse condensin hinge domain highlight common and diverse features of SMC proteins.

Nucleic Acids Research 38, 3454–3465.

Haering, C.H., and Gruber, S. (2016). SnapShot: SMC Protein Complexes Part I. Cell 164,

326–326.

Haering, C.H., Löwe, J., Hochwagen, A., and Nasmyth, K. (2002). Molecular Architecture of

SMC Proteins and the Yeast Cohesin Complex C-terminal domains forming a head would be

part of. Molecular Cell 9, 773–788.

Hassler, M., Shaltiel, I.A., and Haering, C.H. (2018). Towards a Unified Model of SMC

Complex Function. Current Biology 28, R1266–R1281.

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 23: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Hirano, T. (2016). Condensin-Based Chromosome Organization from Bacteria to Vertebrates.

Cell 164, 847–857.

Houlard, M., Godwin, J., Metson, J., Lee, J., Hirano, T., and Nasmyth, K. (2015). Condensin

confers the longitudinal rigidity of chromosomes. NATURE CELL BIOLOGY 17, 771–781.

Hwang, W., and Karplus, M. (2019). Structural basis for power stroke vs. Brownian ratchet

mechanisms of motor proteins. Proceedings of the National Academy of Sciences 116,

19777–19785.

Japaridze, A., Orlandini, E., Smith, K.B., Gmür, L., Valle, F., Micheletti, C., and Dietler, G.

(2017a). Spatial confinement induces hairpins in nicked circular DNA. Nucleic Acids

Research 45, 4905–4914.

Japaridze, A., Muskhelishvili, G., Benedetti, F., Gavriilidou, A.F.M., Zenobi, R., De Los

Rios, P., Longo, G., and Dietler, G. (2017b). Hyperplectonemes: A Higher Order Compact

and Dynamic DNA Self-Organization. Nano Letters 1938−1948.

Jeppsson, K., Kanno, T., Shirahige, K., and Sjögren, C. (2014). The maintenance of

chromosome structure: Positioning and functioning of SMC complexes. Nature Reviews

Molecular Cell Biology 15, 601–614.

Kamada, K., Su’etsugu, M., Takada, H., Miyata, M., and Hirano, T. (2017). Overall Shapes of

the SMC-ScpAB Complex Are Determined by Balance between Constraint and Relaxation of

Its Structural Parts. Structure 25, 603-616.e4.

Katan, A.J., and Dekker, C. (2011). High-Speed AFM Reveals the Dynamics of Single

Biomolecules at the Nanometer Scale. Cell 147, 979–982.

Kim, Y., Shi, Z., Zhang, H., Finkelstein, I.J., and Yu, H. (2019). Human cohesin compacts

DNA by loop extrusion. Science.

Kodera, N., Yamamoto, D., Ishikawa, R., and Ando, T. (2010). Video imaging of walking

myosin V by high-speed atomic force microscopy. Nature 468, 72–76.

Kschonsak, M., Merkel, F., Bisht, S., Metz, J., Rybin, V., Hassler, M., and Haering, C.H.

(2017). Structural Basis for a Safety-Belt Mechanism That Anchors Condensin to

Chromosomes. Cell 171, 588–600.

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 24: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Ku, B., Lim, J.H., Shin, H.C., Shin, S.Y., and Oh, B.H. (2010). Crystal structure of the MukB

hinge domain with coiled-coil stretches and its functional implications. Proteins: Structure,

Function and Bioinformatics 78, 1483–1490.

Kumar, R., Grosbart, M., Nurse, P., Bahng, S., Wyman, C.L., and Marians, K.J. (2017). The

bacterial condensin MukB compacts DNA by sequestering supercoils and stabilizing

topologically isolated loops. Journal of Biological Chemistry 292, 16904–16920.

Li, Y., Schoeffler, A.J., Berger, J.M., and Oakley, M.G. (2010). The Crystal Structure of the

Hinge Domain of the Escherichia coli Structural Maintenance of Chromosomes Protein

MukB. Journal of Molecular Biology 395, 11–19.

Liu, Y., Sung, S., Kim, Y., Li, F., Gwon, G., Jo, A., Kim, A., Kim, T., Song, O., Lee, S.E., et

al. (2016). ATP ‐ dependent DNA binding, unwinding, and resection by the Mre11/Rad50

complex . The EMBO Journal 35, 743–758.

Lyumkis, D. (2019). Challenges and opportunities in cryo-EM single-particle analysis.

Journal of Biological Chemistry 294, 5181–5197.

Marko, J.F., De Los Rios, P., Barducci, A., and Gruber, S. (2019). DNA-segment-capture

model for loop extrusion by structural maintenance of chromosome (SMC) protein

complexes. Nucleic Acids Research 47, 6956–6972.

Moreno-Herrero, F., De Jager, M., Dekker, N.H., Kanaar, R., Wyman, C., and Dekker, C.

(2005). Mesoscale conformational changes in the DNA-repair complex Rad50/Mre11/Nbs1

upon binding DNA. Nature 437, 440–443.

Nasmyth, K., and Haering, C.H. (2005). the Structure and Function of Smc and Kleisin

Complexes. Annual Review of Biochemistry 74, 595–648.

Naumova, N., Imakaev, M., Fudenberg, G., Zhan, Y., Lajoie, B.R., Mirny, L.A., and Dekker,

J. (2013). Organization of the mitotic chromosome. Science 342, 948–953.

Nečas, D., and Klapetek, P. (2012). Gwyddion: An open-source software for SPM data

analysis. Central European Journal of Physics 10, 181–188.

Nichols, M.H., and Corces, V.G. (2018). A tethered-inchworm model of SMC DNA

translocation. Nature Structural & Molecular Biology 25, 906–910.

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 25: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Niki, H. (2017). Open and Closed Questions about Open and Closed SMC. Structure 25, 569–

570.

Noble, A.J., Wei, H., Dandey, V.P., Zhang, Z., Tan, Y.Z., Potter, C.S., and Carragher, B.

(2018). Reducing effects of particle adsorption to the air–water interface in cryo-EM. Nature

Methods 15, 793–795.

Oliveira, R.A., Coelho, P.A., and Sunkel, C.E. (2005). The Condensin I Subunit Barren/CAP-

H Is Essential for the Structural Integrity of Centromeric Heterochromatin during Mitosis †

Downloaded from. MOLECULAR AND CELLULAR BIOLOGY 25, 8971–8984.

Park, Y.B., Hohl, M., Padjasek, M., Jeong, E., Jin, K.S., Krȩzel, A., Petrini, J.H.J., and Cho,

Y. (2017). Eukaryotic Rad50 functions as a rod-shaped dimer. Nature Structural and

Molecular Biology 24, 248–257.

Piazza, I., Rutkowska, A., Ori, A., Walczak, M., Metz, J., Pelechano, V., Beck, M., and

Haering, C.H. (2014). Association of condensin with chromosomes depends on DNA binding

by its HEAT-repeat subunits. Nature Structural & Molecular Biology 21, 560–568.

Rao, S.S.P., Huang, S.C., Glenn St Hilaire, B., Engreitz, J.M., Perez, E.M., Kieffer-Kwon,

K.R., Sanborn, A.L., Johnstone, S.E., Bascom, G.D., Bochkov, I.D., et al. (2017). Cohesin

Loss Eliminates All Loop Domains. Cell 171, 305–320.

Rivetti, C., Guthold, M., and Bustamante, C. (1996). Scanning force microscopy of DNA

deposited onto mica: equilibration versus kinetic trapping studied by statistical polymer chain

analysis. Journal of Molecular Biology 264, 919–932.

Rojowska, A., Lammens, K., Seifert, F.U., Direnberger, C., Feldmann, H., and Hopfner, K.

(2014). Structure of the Rad50 DNA double‐ strand break repair protein in complex with

DNA . The EMBO Journal 33, 2847–2859.

Rowley, M.J., and Corces, V.G. (2018). Organizational principles of 3D genome architecture.

Nature Reviews Genetics 19, 789–800.

van Ruiten, M.S., Rowland, B.D., Ruiten, M.S. Van, and Rowland, B.D. (2018). SMC

Complexes: Universal DNA Looping Machines with Distinct Regulators. Trends in Genetics

34, 477–487.

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 26: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Sanborn, A.L., Rao, S.S.P., Huang, S.-C., Durand, N.C., Huntley, M.H., Jewett, A.I.,

Bochkov, I.D., Chinnappan, D., Cutkosky, A., Li, J., et al. (2015). Chromatin extrusion

explains key features of loop and domain formation in wild-type and engineered genomes.

Proceedings of the National Academy of Sciences 112, E6456–E6465.

Schiener, J., Witt, S., Stark, M., and Guckenberger, R. (2004). Stabilized atomic force

microscopy imaging in liquids using second harmonic of cantilever motion for setpoint

control. Review of Scientific Instruments 75, 2564–2568.

Seifert, F.U., Lammens, K., Stoehr, G., Kessler, B., and Hopfner, K. (2016). Structural

mechanism of ATP ‐ dependent DNA binding and DNA end bridging by eukaryotic Rad50 .

The EMBO Journal 35, 759–772.

Soh, Y.M., Bürmann, F., Shin, H.C., Oda, T., Jin, K.S., Toseland, C.P., Kim, C., Lee, H.,

Kim, S.J., Kong, M.S., et al. (2015). Molecular basis for SMC rod formation and its

dissolution upon DNA binding. Molecular Cell 57, 290–303.

Terakawa, T., Bisht, S., Eeftens, J.M., Dekker, C., Haering, C.H., and Greene, E.C. (2017).

The condensin complex is a mechanochemical motor that translocates along DNA. Science

358, 672–676.

Uchihashi, T., Iino, R., Ando, T., and Noji, H. (2011). High-Speed Atomic Force Microscopy

Reveals Rotary Catalysis of Rotorless F1-ATPase. Science 333, 755–758.

Uchihashi, T., Kodera, N., and Ando, T. (2012). Guide to video recording of structure

dynamics and dynamic processes of proteins by high-speed atomic force microscopy. Nature

Protocols 7, 1193–1206.

Uhlmann, F. (2016). SMC complexes: From DNA to chromosomes. Nature Reviews

Molecular Cell Biology 17, 399–412.

Woo, J.S., Lim, J.H., Shin, H.C., Suh, M.K., Ku, B., Lee, K.H., Joo, K., Robinson, H., Lee, J.,

Park, S.Y., et al. (2009). Structural Studies of a Bacterial Condensin Complex Reveal ATP-

Dependent Disruption of Intersubunit Interactions. Cell 136, 85–96.

Yoshimura, S.H., Hizume, K., Murakami, A., Sutani, T., Takeyasu, K., and Yanagida, M.

(2002). Condensin Architecture and Interaction with DNA. Current Biology 12, 508–513.

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 27: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Zhang, N., Kuznetsov, S.G., Sharan, S.K., Li, K., Rao, P.H., and Pati, D. (2008). A handcuff

model for the cohesin complex. Journal of Cell Biology 183, 1019–1031.

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 28: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Figure 1. The condensin holocomplex exhibits different shapes.

(A) Cartoon of the yeast condensin holocomplex. (B) Typical dry AFM image of condensin. (C)

Images of various conformations of condensin holocomplexes (N = 384). Small panels are dry-

AFM images and large panels are liquid-AFM images of the O- and B-shape. In the AFM image

of kleisin-opened state, the green arrow indicates the non-SMC domains while the orange arrow

points to an smc end. In the AFM image of the hinge-opened state, orange arrows indicate the

dissociated hinge domains. (D) Relative occurrence of different states in the presence of ATP. (E)

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 29: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Distance between hinge and the globular domain for O shape, kleisin-opened shape, and I-shape

condensins. (F) Hinge angle measured at the hinge domain. * P<0.05, ** P<0.01, and *** P<0.001,

assessed using the paired t test. (G) Relative occurrence of various conformations (N=384 without

ATP, N=255 with ATP, and N=419 for the EQ mutant with ATP). Error bars indicate SD unless

specified otherwise. 

   

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 30: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Figure 2. Condensin dynamically toggles between an O-shape and B-shape over time.

(A) Representative HS liquid AFM images of the condensin holocomplex followed over time at a

5 frames/s rate (Movie S1). The complex is seen to change its conformation over time, from an O

shape (62-63.5 s) to a collapsed B-shape (after 63.5-64 s). (B) Real-time trace of the hinge-head

distance. The distance is seen to toggle between a large and small value, corresponding to the O-

and B-shapes respectively. The distribution of the hinge-head distance, depicted on the right,

clearly shows two peaks. Solid line is a fit of two Gaussians. (C) Average peak-to-peak distance of

17 ± 7 nm from many distributions such as in panel B (N=12 individual traces taken on different

SMC molecules).

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 31: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Figure 3. Binding of condensin to DNA.

(A) Binding of condensin to DNA through its globular domain, presumably through the Heat

domains. (B) Binding of condensin to DNA through its hinge. (C) Binding of condensin to DNA

at two sites. (D) Collapsed B-shape condensin on DNA. (E) Relative occurrence of each of the

different binding modes (N=262).

 

 

 

 

 

 

 

 

 

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 32: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Figure 4. Condensin-mediated DNA loops.

(A-B) Representative condensin-mediated DNA loops imaged using liquid AFM (A) and dry AFM

(B). Cartoons are a guide to the eye. (C) Number of DNA loops per µm of DNA under various

conditions: with ATP (150 loops), with ATPγS (21 loops), without ATP (24 loops), and without

Ycg1 (15 loops), as measured on N ≥5 independent experiments. Loop formation is strongly

dependent on ATP hydrolysis and on the presence of the Ycg1 domain. (D) Loop size comparison

in various conditions (median ± SEM). (E) Distribution of the number of condensins found at the

stem of a DNA loop (N=96).

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 33: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Figure 5. The neck size of the DNA loop correlates with the size of the condensin complex.

(A) Top left schematic indicates the protein width and neck size of the DNA loop. Lines in the

image on the top right indicate where crosssectional height profiles are acquired, with the result

denoted at the bottom panel. (B) Neck size of DNA loops versus width of the condensin at the loop

stem. Every point comes from a measurement of a single loop (N=115). The distribution of neck

sizes (shown on the right) is fitted to two Gaussian, yielding average peak values of 9 ± 1 and 31

± 20 nm.

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint

Page 34: AFM images of open and collapsed states of yeast condensin … · 2019. 12. 13. · two HEAT-repeat subunits (Ycg1 and Ycs4), and the Brn1 kleisin subunit. In dry AFM, these could

Figure 6. Working model for condensin-mediated DNA loop extrusion.

Condensin binds DNA tightly to the Ycg1 anchor, as well as, transiently, to two other binding sites.

From an open O-shape (A) it collapses into a B-shape (B) where the neck of the DNA loop

concurrently changes from wide to narrow. Subsequently, the hinge is released (C), and the hinge

reaches out to bind new DNA (D). (E) Upon repeating this cycle, DNA is extruded into a loop.

preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission. The copyright holder for thisthis version posted December 13, 2019. ; https://doi.org/10.1101/2019.12.13.867358doi: bioRxiv preprint