<<

1 Compaction and segregation of sister

2 via active loop extrusion

3

4

5 Anton Goloborodko1, Maxim V. Imakaev1, John F. Marko2, and Leonid Mirny1,3,*

6

7 Short title: Compaction and segregation of sister chromatids

8

9

10 1 Department of Physics, Massachusetts Institute of Technology, Cambridge MA 02139

11 2 Department of Molecular Biosciences and Department of Physics and Astronomy,

12 Northwestern University, Evanston IL 60208

13 3 Institute for Medical Engineering & Science, Massachusetts Institute of Technology,

14 Cambridge MA 02139

15

16 * Corresponding author: Leonid Mirny, Institute for Medical Engineering & Science,

17 Massachusetts Institute of Technology, 77 Mass Ave, Cambridge MA 02139, [email protected],

18 +16174524862

19 Keywords: , polymers, condensin, structural maintenance of

20

1 21

22 Abstract

23

24 The mechanism by which chromatids and chromosomes are segregated during

25 and is a major puzzle of biology and biophysics. Using polymer simulations of

26 dynamics, we show that a single mechanism of loop extrusion by

27 condensins can robustly compact, segregate and disentangle chromosomes, arriving at

28 individualized chromatids with morphology observed in vivo. Our model resolves the

29 paradox of topological simplification concomitant with chromosome “condensation”, and

30 explains how enzymes a few nanometers in size are able to control chromosome

31 geometry and topology at micron length scales. We suggest that loop extrusion is a

32 universal mechanism of genome folding that mediates functional interactions during

33 and compacts chromosomes during mitosis.

34

2 35 Introduction

36 The mechanism whereby eukaryote chromosomes are compacted and concomitantly

37 segregated from one another remains poorly understood. A number of aspects of this

38 process are remarkable. First, the chromosomes are condensed into elongated

39 structures that maintain the linear order, i.e. the order of genomic elements in the

40 elongated chromosome resembles their order along the genome 1. Second, the

41 compaction machinery is able to distinguish different chromosomes and chromatids,

42 preferentially forming intra- cross-links: if this were not the case, segregation

43 would not occur 2. Third, the process of compaction is coincident with segregation of

44 sister chromatids, i.e. formation of two separate chromosomal bodies. Finally, originally

45 intertwined sister chromatids become topologically disentangled, which is surprising,

46 given the general tendency of polymers to become more intertwined as they are

47 concentrated 3.

48

49 None of these features cannot be produced by indiscriminate cross-linking of

50 chromosomes 4, which suggests that a novel mechanism of polymer compaction must

51 occur, namely “lengthwise compaction” 3,5,6, which permits each chromatid to be

52 compacted while avoiding sticking of separate chromatids together. Cell-biological

53 studies suggest that topoisomerase II and condensin are essential for metaphase

54 chromosome compaction 7–10, leading to the hypothesis that mitotic compaction-

55 segregation relies on the interplay between the activities of these two protein

56 complexes. Final structures of mitotic chromosomes were shown to consist arrays of

57 consecutive loops 11–13. Formation of such arrays would naturally results in lengthwise

58 chromosome compaction.

59

3 60 One hypothesis of how condensins can generate compaction without crosslinking of

61 topologically distinct chromosomes is that they bind to two nearby points and then slide

62 to generate a progressively larger loop 2. This “loop extrusion” process creates an

63 array of consecutive loops in individual chromosomes 2. When loop-extruding

64 condensins exchange with the solvent, the process eventually settles at a dynamical

65 steady state with a well-defined average loop size14. Simulations of this system at larger

66 scales 14 have established that there are two regimes of the steady-state dynamics: (i) a

67 sparse regime where little compaction is achieved and, (ii) a dense regime where a

68 dense array of stabilized loops efficiently compacts a chromosome. Importantly, loop

69 extrusion in the dense regime generates chromatin loops that are stabilized by multiple

70 “stacked” condensins 15, making loops robust against the known dynamic binding-

71 unbinding of individual condensin complexes 16., These two quantitative studies of loop

72 extrusion kinetics14,15 focused on the hierarchy of extruded loops and did not consider

73 the 3D conformation and topology of the chromatin fiber in the formed loop arrays. The

74 question of whether loop-extruding factors can act on a chromatin fiber so as to form an

75 array of loops, driving chromosome compaction and chromatid segregation, as originally

76 hypothesized for condensin in 2, is salient and unanswered. The main objective of this

77 paper is to test this hypothesis and to understand how formation of extruded loop arrays

78 ultimately leads to compaction, segregation and disentanglement (topology

79 simplification) of originally intertwined sister chromatids.

80

81 Here we use large-scale polymer simulations to show that active loop extrusion in

82 presence of topo II is sufficient to reproduce robust lengthwise compaction of chromatin

83 into dense, elongated, prophase chromatids with morphology in quantitatively accord

84 with experimental observations. Condensin-driven lengthwise compaction, combined

85 with the strand passing activity of topo II drives disentanglement and segregation of 4 86 sister chromatids in agreement with theoretical prediction that linearly compacted

87 chromatids must spontaneously disentangle 3.

88

89

90 Model

91 We consider a chromosome as a flexible polymer, coarse-graining to monomers of 10

92 nm diameter, each corresponding to three nucleosomes (600bp). As earlier 13,17, the

93 polymer has a persistence length of ~5 monomers (3Kb), is subject to excluded volume

94 interactions and to the activity of loop-extruding condensin molecules and topo II (see

95 below, and in 17). We simulate chains of 50000 monomers, which corresponds to 30 Mb,

96 close to the size of the smallest human chromosomal arm.

97

98 Each condensin complex is modeled as a dynamic bond between a pair of monomers

99 that is changed as a function of time (Figure 1, Video 1). Upon binding, each

100 condensin forms a bond between two adjacent monomers; subsequently both bond

101 ends of a condensin move along the chromosome in opposite directions, progressively

102 bridging more distant sites and effectively extruding a loop. As in prior lattice models of

103 condensins 14,15, when two condensins collide on the chromatin, their translocation is

104 blocked; equivalently, only one condensin is permitted to bind to each monomer,

105 modeling their steric exclusion. Exchange of condensins between chromatin and

106 solution is modeled by allowing each condensin molecule to stochastically unbind from

107 the polymer. To maintain a constant number of bound condensins, when one

108 condensin molecule dissociates, another molecule associates at a randomly chosen

109 location, including chromatin within loops extruded by other condensins. This can

110 potentially lead to formation of reinforced loops (Video 2, see below) 14,15.

5 111

112 To simulate the strand-passing activity of topo II enzymes, we permit crossings of

113 chromosomal fiber by setting a finite energy cost of fiber overlaps. By adjusting the

114 overlap energy cost we can control the rate at which topology changes occur.

115

116 This model has seven parameters. Three describe the properties of the chromatin fiber

117 (linear density, fiber diameter, and persistence length); two control condensin kinetics

118 (the mean linear separation between condensins, and their processivity, i.e. the average

119 size of a loop formed by an isolated condensin before it dissociates, and defined as two

120 times the velocity of bond translocation times the mean residence time); one parameter

121 controls the length of the monomer-monomer bond mediated by a condensin (effectively

122 the size of a condensin complex), and finally we have the energy barrier to topological

123 change.

124

125 The effects of two of these parameters, linear separation of condensins and their

126 processivity, have been considered earlier in the context of a 1D model 14. These

127 parameters control the fraction of a chromosome extruded into loops and the average

128 loop length. In our simulations, we use the separation of 30kb (1000 condensins per 30

129 Mb chromosome, as measured in 18) and the processivity of 830 kb such that

130 condensins form a dense array of gapless loops with the average loop length of 100kb,

131 which agrees with available experimental observations 11–13,19. The effects of the other

132 five parameters are considered below.

133

134 Initially, chromosomes are compacted to the density of chromatin inside the human

135 nucleus and topologically equilibrated, thus assuming an equilibrium globular

136 conformation, corresponding to a chromosomal territory. In the simulations of sister 6 137 chromatid segregation, the initial conformation of one chromatid is generated as above,

138 while the second chromatid traces the same path at a distance of 50nm and winds one

139 full turn per 100 monomers (60 kb). In this state, two chromatids run parallel to each

140 other and are highly entangled, while residing in the same chromosome territory. This

141 represents to the most challenging case for segregation of sister chromatids.

142

143

144 Results

145 Loop extrusion compacts randomly coiled interphase chromosomes into

146 prophase-like chromosomes

147

148 Our main qualitative result is that loop extrusion by condensins and strand passing

149 mediated by topo II convert initially globular interphase chromosomes into elongated,

150 unknotted and relatively stiff structures that closely resemble prophase chromosomes

151 (Figure 2, Video 2, Figure 2 - figure supplement 1). The compacted chromosomes

152 have the familiar “noodle” morphology (Figure 2BC) of prophase chromosomes

153 observed in optical and electron microscopy 20,21. Consistent with the observed linear

154 organization of prophase chromosomes, our simulated chromosomes preserve the

155 underlying linear order of the genome20–24 (Figure 2C). Moreover, the geometric

156 parameters of our simulated chromosomes match the experimentally measured

157 parameters of mid-prophase chromosomes (Figure 2B): both simulated and human

158 chromosomes have a linear compaction density ~5 kb/nm 20 and the radius of 300nm

159 25–27.

160

161 The internal structure of compacted chromosomes also agrees with structural data for

7 162 mitotic chromosomes. First, loop extrusion compacts a chromosome into a chain of

163 consecutive loops 14, which agrees with a wealth of microscopy observations 11,12,28 and

164 the recent Hi-C study 13. Second, our polymer simulations show that a chain of loops

165 naturally assumes a cylindrical shape, with loops forming the periphery of the cylinder

166 and loop-extruding condensins at loop bases forming the core (Figure 2C, Video 2), as

167 was predicted in 2. Condensin-staining experiments reveal similar cores in the middle of

168 in vivo and reconstituted mitotic chromosomes 28,29. In agreement with experiments

169 where condensin cores appear as early as late prophase 27, we observe rapid formation

170 of condensin core in simulations (Figure 2, Video 2). Thus, the structure of

171 chromosomes compacted by loop extrusion is consistent with the loop-compaction

172 picture of mitotic 11,29,30 and meiotic chromosomes 31,32. However, it should be noted

173 that our model does not rely on a connected “scaffold”, and that cleavage of the DNA

174 will result in disintegration of the entire structure, as is observed experimentally 33.

175

176 Loop-extrusion generates chromosome morphology kinetics similar to those

177 observed experimentally

178 Simulations also reproduce several aspects of compaction kinetics observed

179 experimentally. First, the model shows that formed loops are stably maintained by

180 condensins despite their constant exchange with the nucleoplasm 8,16,34. This stability is

181 achieved by accumulation of several condensins at a loop base 14,15. Supported by

182 multiple condensins, each loop persists for times much longer than the residency time

183 for individual condensins 14.

184

185 Second, simulations show that chromosome compaction proceeds by two stages

186 (Figure 2AB): fast formation of a loose, elongated morphology, followed by a slow

187 maturation stage. During the initial “loop formation stage”, condensins associate with 8 188 the chromosome and extrude loops until colliding with each other, rapidly compacting

189 the chromosome (Figure 2, Video 2). At the end of this stage, that takes a fraction of

190 the condensin exchange time, a gapless array of loops is formed, the loops are relative

191 small, supported by single condensins, and the chromosome is long and thin (Figure

192 2B).

193

194 During the following “maturation stage”, due to condensin exchange some loops

195 dissolve and divide while others grow and get reinforced by multiple condensins14,15. As

196 we recently showed14, a loop can dissolve when all of its condensins dissociate; while

197 new loops are born when two condensins land inside an existing loop at about the same

198 time and split it into two loops. As the mean loop size grows during the maturation stage

199 (Figure 2A), the chromosome becomes thicker and shorter. By the end this stage,

200 which takes several rounds of condensin exchange, the rates of loop death becomes

201 equal the rate of loop division and the chromosome achieves a steady state. At steady

202 state loops get reinforced by multiple condensins and the average loop size and the

203 lengthwise compaction reach their maximal values (Figure 2). These dynamics are in

204 accord with the observation that chromosomes rapidly attain a condensed “noodle”-like

205 shape in early prophase, then spend the rest of prophase growing thicker and shorter

206 20,21,27,35 (see Discussion).

207

208 Loop extrusion separates and disentangles sister chromatids

209 In a second set of simulations, we studied whether loop extrusion and strand passing

210 simultaneously lead to spatial segregation 2 and disentanglement of sister chromatids 5.

211 We simulated two long polymers connected at their midpoints that represented sister

212 chromatids with stable centromeric cohesion. To model cohesion of sister chromatids in

213 late we further twisted them around each other (see Methods) while 9 214 maintaining them with a chromosomal territory.

215

216 We find that the activity of loop-extruding condensins in the presence of topo II leads to

217 compactions of each of the sister chromatids into prophase-like “noodle” structures

218 (Video 3). Moreover, we observe that formed compact chromosomes (a) become

219 segregated from each other, as evident by the drastic growth of the mean distance

220 between their backbones (Figure 3A), and (b) become disentangled, as shown by the

221 reduction of the winding number of their backbones (Figure 3B). Thus, loop extrusion

222 and strand passing folds highly catenated sister chromatids into separate disentangled

223 structures. (Figure 3C, Video 3).

224

225 Despite stochastic dynamics of condensins, compaction, segregation and

226 disentanglement of sister chromatids are highly reproducible and robust to changes in

227 simulation parameters. Disentanglement and segregation were observed in each of 16

228 simulation replicas (Video 4). We also performed simulations with altered chromatin

229 polymer parameters (Video 5). Again, we observed chromatids segregation and

230 disentanglement irrespective of fiber parameters and size of a single condensin

231 molecule.

232

233 Our model makes several predictions that can be tested against available experimental

234 data. First, in agreement with the published literature36–39 the abundance of condensins

235 on the chromosome significantly affects geometry of simulated chromatids (Video 6).

236 When we varied abundance of chromosome-bound condensin by 20-fold (5-fold

237 decrease and 4-fold increase) we observed two major effects (Figure 2 - figure

238 supplement 1). First, we noticed that condensin depletion leads to a systematic (3-fold)

239 increase in chromosome diameter. Second, we observed a nonmonotonic changes in 10 240 the linear density of chromatin in compacted chromosomes: both 5-fold increase and 4-

241 fold decrease of condensin abundance lead to about 2-fold drop in linear compaction

242 (i.e. 2-fold increase in chromosome length). This reduced compaction, however,

243 emerges due to different mechanisms: an increase of condensin abundance leads to

244 reduction of loop sizes while keeping all chromatin in the loops, creating thin

245 chromosomes with long scaffold formed by excessive condensins. Decrease of

246 condensins abundance, on the contrary, leads to formation of chromosomes where

247 some regions are not extruded into loops; such gaps between loops lead to substantial

248 increase in chromosome length and inefficient compaction. These nontrivial

249 dependences of chromosome geometry on condensin abundance can be further tested

250 experimentally.

251

252 Next, we examined the role of topological constraints and topo II enzyme in the process

253 of chromosome compaction and segregation. Activity of topo II is modeled by a finite

254 energy barrier for monomer overlap, allowing for occasional fiber crossings. We

255 simulated topo II depletion by elevating the energy barrier and increasing the radius of

256 repulsion between distant particles, thus reducing the rate of fiber crossings. In these

257 simulations segregation of sister chromatids is drastically slowed down (Video 7). This

258 effect of topological constraints on kinetics of compaction and segregation is expected

259 since each chromosome is a chain with open ends that thus can eventually segregate

260 from each other. Reduction of topo II activity (increase of the crossing barrier) prevents

261 disentanglement and segregation of sister chromatids, while having little effect of

262 diameter of individual chromatids (Figure 3 – figure supplement 1). We conclude that

263 topo II is essential for rapid chromosome segregation and disentanglement, in accord

264 with an earlier theoretical work 40 and with experimental observations 9,41–43.

265 11 266 Finally, we examined the role of -dependent cohesion of chromosomal arms of

267 sister chromatids. It is known that are actively removed from chromosomal

268 arms in prophase. Interference with processes of cohesins removal or cleavage leads to

269 compaction into elongated chromosomes that fail to segregate 44,45. We simulated

270 prophase compaction in the presence of cohesin rings 46 that stay on sister chromatids

271 and can be pushed around by loop-extruding condensins. If cohesins remain uncleaved

272 on sister chromatids (Video 8), we observe compaction of sister chromatids, that

273 remain unsegregated, in agreement with experiment 44,45.

274

275 Taken together, these results indicate that loop extrusion is sufficient to compact,

276 disentangle and segregated chromosomes from a more open interphase state, to a

277 compacted, elongated and linearly organized prophase state. Future Hi-C experiments

278 focused on prophase chromosomes may be able to make detailed structural

279 comparison with the prediction of our model.

280

281

282 Repulsion between loops shapes and segregates compacted chromatids.

283 Reorganization of chromosomes by loop extrusion can be rationalized by considering

284 physical interactions between extruded loops. A dense array of extruded loops makes a

285 chromosome acquire a bottle-brush organization with loops forming side chains of the

286 brush. Polymer bottle-brushes has been studied in physics both experimentally and

287 computationally 47–49, and has been proposed as models of mitotic and meiotic

288 chromosomes 4. A recent study also suggested a bottle-brush organization of

289 in yeast, where chromatin loops lead to separation of sister

290 centromeres and create spring-like properties of the centromere in mitosis 50,51.

291 12 292 Central to properties of chromosomal bottle-brushes are excluded volume interactions

293 and maximization of conformational entropy that lead to repulsion of loops from each

294 other 4,52. Loops further stretch from the core and extend radially (Figure 4), similar to

295 stretching of side chains in polymer bottle-brushes 47–49. This in turn leads to formation

296 of the central core where loop-bases accumulate (Figure 4A). Strong steric repulsion

297 between the loops, caused by excluded volume interactions, stretches the core, thus

298 maintaining linear ordering of loop bases along the genome (Figure 2C). Repulsion

299 between loops also leads to stiffening of the formed loop-bottle-brush, as chromosome

300 bending would increase overlaps between the loops.

301

302 Steric repulsion between loops of sister chromatid leads to segregation of sister

303 chromatids that minimizes the overlap of their brush coronas (Figure 4B). Stiffening of

304 the bottle-brush, in turn, leads to disentanglement of sister chromatids. To test this

305 mechanism we performed simulations where excluded volume interactions have been

306 turned off (Video 5). In these simulations, chromosomes did not assume a “bottle-

307 brush” morphology and did not segregate, which confirms the fundamental role of steric

308 repulsion for chromosome segregation.

309

310 Discussion

311 Our results show that chromosomal compaction by loop extrusion generates the major

312 features of chromosome reorganization observed during prophase 2,15. Loop extrusion

313 leads to lengthwise compaction and formation of cylindrical, brush-like chromatids,

314 reducing chromosome length by roughly 25-fold, while at the same time removing

315 entanglements between separate chromatids. Unlike models of prophase condensation

316 based on crosslinking of randomly colliding fibers 52,53, chromatid compaction by loop

13 317 extrusion exclusively forms bridges between loci of the same chromosome (Figure 1A).

318 Similarly, loop-extrusion in bacteria, mediated by SMC proteins, was suggested to form

319 intra-arm chromosomal bridges leading to lengthwise chromosome compaction 54,55.

320

321 Our computations indicate that it takes an extended period of time for loop-extruding

322 enzymes to maximally compact chromosomes. Maximal compaction is achieved by

323 slow maturation of the loop array over multiple rounds of condensin exchange, rather

324 than a single round of loop expansion. Loop length and the number of condensins per

325 loop gradually increase during the maturation stage. Assuming that condensin II in

326 prophase exchanges with the fast rate of condensin I in metaphase (~200 sec) 16 and

327 using an average loop size of 100 kb, and an average spacing of condensins of 30 kb, it

328 should take ~5 condensin exchange times, or about 17 minutes to achieve maximal

329 compaction, which roughly agrees with the duration of prophase in human cells 56,57.

330 This gradual compaction also may explain why chromosomes of cells arrested in

331 metaphase continue shrinking for many hours after the arrest (condensin II metaphase

332 exchange time ~hours 16).

333

334 Our model is idealized and does not aim to describe all of aspects of chromosome

335 folding during mitosis. First, our model does not fully capture the specifics of later

336 stages of mitosis. We model only the action of condensin II in prophase 58. Simple

337 geometric considerations, Hi-C studies and mechanical perturbation of mitotic

338 chromosomes show that, in metaphase, cells use additional mechanisms of lengthwise

339 (axial) compaction.

340

341 Loop extrusion does not require specific interaction sites (e.g. loop anchors or

342 condensin binding sites) and is therefore robust to mutations and chromosome 14 343 rearrangements. This robustness also explains how a large segment of chromatin

344 inserted into a chromosome of a different species can be reliably compacted by mitotic

345 machinery of the host 59,60. At the same time, this model allows the morphology of

346 mitotic chromosomes to be tuned. Most importantly, the diameter and length of the

347 simulated chromosomes depend on chromatin fiber properties as well as numbers of

348 condensins (Figure 2 - figure supplement 1); those dependences invite experimental

349 tests.

350

351 In summary, we have shown that loop extrusion is a highly robust mechanism to

352 compact, segregate and disentangle chromosomes. In a recent study 17 we also

353 demonstrated that loop extrusion by another SMC complex, cohesin, could lead to

354 formation of interphase topological association domains (TADs). We suggest that during

355 interphase, cohesins are sparsely bound, extruding 50-70% of DNA into highly dynamic

356 loops, while metaphase condensins are bound more densely and more processive

357 forming a dense array of stable loops. Taken together these studies suggest that loop

358 extrusion can be a universal mechanism of genome folding, to mediate functional

359 interactions during interphase and to compact during mitosis.

360

15 361 Materials and methods

362 Methods

363 Our model of loop extrusion acting on a polymer fiber consists of (a) a 1D model that

364 governs the dynamics of intra-chromosomal bonds formed by condensins over time and

365 (b) a 3D polymer model of chromosome dynamics subject to the condensins bonds

366 The 1D model of loop extrusion

367 The dynamics of loop extruding condensins on a chromatin fiber is simulated using a 1D

368 lattice simulations as it was described previously for generic loop-extruding factors

369 14,15 In the lattice, each position corresponds to one monomer in the 3D polymer

370 simulation. We model a condensin molecule very generally as having two chromatin

371 binding sites or, “heads”, connected by a linker. Each head of a condensin occupies one

372 lattice position at a time, and no two heads can occupy the same lattice position. To

373 simulate the process of loop extrusion, the positions of the two condensin heads

374 stochastically move away from each other over time (simulated with the Gillespie

375 algorithm 61).

376

377 We initialize the simulations by placing condensin molecules at random positions along

378 the polymer chain, with both heads in adjacent positions. To simulate the exchange of

379 condensins between chromatin and solution, condensins stochastically dissociate from

380 the chromatin fiber. Every dissociation event is immediately followed by association of

381 another condensin molecule with the chromatin fiber so that the total number of

382 condensins bound to the chromatin stays constant.

383

384 This 1D model has four parameters: the size of the lattice, number of condensins bound

385 to chromatin, speed of extrusion, average residency time of a condensin on the

16 386 chromatin fiber. We model a 30 Mb chromosome with a lattice of 50000 sites, 600 bp

387 each. The chromosome is bound by 1000 condensins. Without loss of generality, we set

388 the speed of extrusion to be 1 step per unit time and the condensin residency time at

389 692 units of time, such that the resulting average loop length is equal 167 monomers,

390 or, 100kb, close to previous observations in vivo.

391

392 3D simulations of chromosomes

393 To perform Langevin dynamics polymer simulations we use OpenMM, a high-

394 performance GPU-assisted molecular dynamics API 62,63. To represent chromatin fibers

395 as polymers, we use a sequence of spherical monomers of 1 unit of length in diameter.

396 Here and below all distances are measured in monomer sizes (~3 nucleosomes,

397 ~10nm), density is measured in particles per cubic unit, and energies are measured in

398 kT. We use the following parameters of the Langevin integrator: particle mass = 1 amu,

399 friction coefficient = 0.01 ps^-1, time step = 1ps, temperature = 300K.

400

401 Neighboring monomers are connected by harmonic bonds, with a potential =

402 100( − 1) (here and below in units of kT). We model polymer stiffness with a three

403 point interaction term, with the potential =5 (1−cos()), where alpha is the angle

404 between neighboring bonds.

405

406 To allow chain passing, which represents activity of topoisomerase II, we use a soft-

407 core potential for interactions between monomers, similar to 13,64. All monomers interact

408 via a repulsive potential

= 2.0(−1 + (/√(6 ⁄ 7))^12 ⋅ ((/√(6 ⁄ 7))^2 − 1) ⋅ (823543 )/46656)

409

17 410 This is a fast and efficient potential designed to be constant at 2.0 kT up to r=0.7-0.8

411 and then quickly go to zero at r=1.00.

412

413 To connect 1D LEF simulations with 3D polymer simulations, we first run 1D LEF

414 dynamics for a total period of 10 condensin residence times, recording the state of the

415 systems each unit of time. We then assign bonds to the monomers in polymer

416 simulations according to the current position of condensins’ heads. The two monomers

417 held by the two heads of each condensin are connected by a harmonic bond with the

418 potential = 100( − 1) . For each position of condensins’ heads from the 1D model,

419 we perform 40000 steps of Langevin dynamics. 3D conformations are recorded every

420 20000 steps.

421

422 We allow an overlap of the heads of collided condensins at the loop bases (i.e. two

423 heads of collided condensins could occupy the same monomer instead of two adjacent

424 monomers). We implement this by shifting the positions of the downstream heads of all

425 condensins by 1 monomer downstream. This allows us to achieve the maximally

426 possible compaction of the chromosomal core (one loop per one monomer of the axis).

427

428 We generate the initial conformation of a single chromosome as following: a polymer

429 chain was spherically compacted to a density of 0.01 particle per unit length^3, then

430 allowed to equilibrate over 4,000,000 steps of Langevin dynamics, with a gradual

431 increase of repulsion energy to equilibrate both the topology and the distribution of

432 density inside the confining sphere. We generate the initial conformations of two sister

433 chromatids by winding of two polymer chains along the conformation of a single

434 chromosome, with one full turn each 100 monomers.

435 18 436 We simulate topo II depletion by adjusting two factors. First, we increase the energy of

437 monomer overlap to 20 kT. Since this measure alone proved to be inefficient to prevent

438 chain passing, we additionally increase the radius of repulsion up to 3 length units. In

439 order to maintain the contour length of the polymer, we keep the length of a monomer

440 bond at 1 and ignore repulsion between pairs of neighboring monomers, up to 3

441 monomers distance along the chain.

442

443 To study how the parameters of simulations affect the geometry of compacted

444 chromosomes, we alter the following parameters: (a) increase 2x and (b) decrease 0.5x

445 the number of condensins, (c) increase 2x and (d) decrease 0.5x the bending energy,

446 (e) disallow overlaps of condensins at loop bases, thus simulating wide condensins; (f)

447 reduce linear DNA density of the chromatin fiber to 400 bp/10nm to model compaction

448 of a 10nm fiber of stacked nucleosomes and (e) increase linear DNA density of the

449 chromatin fiber to 2400 bp/10nm and increased the fiber thickness to 30nm to model

450 compaction of a 30nm fiber.

451

452 Acknowledgements

453 Work at NU was supported by the NSF through Grants DMR-1206868 and MCB-

454 1022117, and by the NIH through Grants GM105847 and CA193419. Work at MIT was

455 supported by the NIH through Grants GM114190, R01HG003143, and DK107980. We

456 thank Elnaz Alipour, Job Dekker, Nancy Kleckner, the MIT Biophysics students and the

457 members of Mirny Lab: Geoff Fudenberg, Nezar Abdennur, Christopher McFarland and,

458 especially, E. M. Breville for stimulating and productive discussions.

459

460

19 461 References

462 (1) Trask, B. J.; Allen, S.; Massa, H.; Fertitta, A. Studies of Metaphase and

463 Interphase Chromosomes Using Fluorescence In Situ Hybridization. Cold Spring

464 Harb Symp Quant Biol 1993, 58, 767–775 DOI: 10.1101/SQB.1993.058.01.084.

465 (2) Nasmyth, K. Disseminating the genome: joining, resolving, and separating sister

466 chromatids during mitosis and meiosis. Annu. Rev. Genet. 2001, 35, 673–745

467 DOI: 10.1146/annurev.genet.35.102401.091334.

468 (3) Marko, J. F. Scaling of Linking and Writhing Numbers for Spherically Confined

469 and Topologically Equilibrated Flexible Polymers. J. Stat. Phys. 2011, 142 (6),

470 1353–1370 DOI: 10.1007/s10955-011-0172-4.

471 (4) Marko, J. F.; Siggia, E. D. Polymer Models of Meiotic and Mitotic Chromosomes.

472 Mol. Biol. Cell 1997, 8 (11), 2217–2231 DOI: 10.1091/mbc.8.11.2217.

473 (5) Marko, J. F. Linking topology of tethered polymer rings with applications to

474 chromosome segregation and estimation of the knotting length. Phys. Rev. E

475 2009, 79 (5), 051905 DOI: 10.1103/PhysRevE.79.051905.

476 (6) Marko, J. F. The Mitotic Chromosome : Structure and Mechanics. In Genome

477 Organization and Function in the Cell Nucleus; Rippe, K., Ed.; Wiley-VCH Verlag

478 GmbH & Co. KGaA, 2011; pp 449–485.

479 (7) Hirano, T.; Mitchison, T. J. Topoisomerase II does not play a scaffolding role in

480 the organization of mitotic chromosomes assembled in Xenopus egg extracts. J.

481 Cell Biol. 1993, 120 (3), 601–612 DOI: 10.1083/jcb.120.3.601.

482 (8) Hirano, T.; Mitchison, T. J. A heterodimeric coiled-coil protein required for mitotic

483 chromosome condensation in vitro. Cell 1994, 79 (3), 449–458 DOI:

484 10.1016/0092-8674(94)90254-2.

485 (9) Wood, E. R.; Earnshaw, W. C. Mitotic chromatin condensation in vitro using

20 486 somatic cell extracts and nuclei with variable levels of endogenous topoisomerase

487 II. J. Cell Biol. 1990, 111 (6), 2839–2850 DOI: 10.1083/jcb.111.6.2839.

488 (10) Hirano, T. Biochemical and genetic dissection of mitotic chromosome

489 condensation. Trends Biochem. Sci. 1995, 20 (9), 357–361 DOI: 10.1016/S0968-

490 0004(00)89076-3.

491 (11) Paulson, J. R.; Laemmli, U. K. The structure of histone-depleted metaphase

492 chromosomes. Cell 1977, 12 (3), 817–828 DOI: 10.1016/0092-8674(77)90280-X.

493 (12) Earnshaw, W. C.; Laemmli, U. K. Architecture of metaphase chromosomes and

494 chromosome scaffolds. J. Cell Biol. 1983, 96 (1), 84–93 DOI: 10.1083/jcb.96.1.84.

495 (13) Naumova, N.; Imakaev, M.; Fudenberg, G.; Zhan, Y.; Lajoie, B. R.; Mirny, L. a;

496 Dekker, J. Organization of the mitotic chromosome. Science 2013, 342 (6161),

497 948–953 DOI: 10.1126/science.1236083.

498 (14) Goloborodko, A.; Marko, J. F.; Mirny, L. Mitotic chromosome compaction via

499 active loop extrusion. bioRxiv 2015 DOI: 10.1101/021642.

500 (15) Alipour, E.; Marko, J. F. Self-organization of domain structures by DNA-loop-

501 extruding enzymes. Nucleic Acids Res. 2012, 40 (22), 11202–11212 DOI:

502 10.1093/nar/gks925.

503 (16) Gerlich, D.; Hirota, T.; Koch, B.; Peters, J.-M.; Ellenberg, J. Condensin I stabilizes

504 chromosomes mechanically through a dynamic interaction in live cells. Curr. Biol.

505 2006, 16 (4), 333–344 DOI: 10.1016/j.cub.2005.12.040.

506 (17) Fudenberg, G.; Imakaev, M.; Lu, C.; Goloborodko, A.; Abdennur, N.; Mirny, L. A.

507 Formation of Chromosomal Domains by Loop Extrusion. bioRxiv 2015 DOI:

508 10.1101/024620.

509 (18) Fukui, K.; Uchiyama, S. Chromosome protein framework from proteome analysis

510 of isolated human metaphase chromosomes. Chem. Rec. 2007, 7 (4), 230–237

511 DOI: 10.1002/tcr.20120. 21 512 (19) Jackson, D.; Dickinson, P.; Cook, P. R. The size of chromatin loops in HeLa cells.

513 EMBO J. 1990, 9 (2), 567–571.

514 (20) Yunis, J. J. Mid-prophase human chromosomes. The attainment of 2000 bands.

515 Hum. Genet. 1981, 56 (3), 293–298 DOI: 10.1007/BF00274682.

516 (21) Sumner, A. T. Scanning electron microscopy of mammalian chromosomes from

517 prophase to telophase. Chromosoma 1991, 100 (6), 410–418 DOI:

518 10.1007/BF00337519.

519 (22) Yunis, J. J.; Sanchez, O. The G-banded prophase chromosomes of man. Hum.

520 Genet. 1975, 27 (3), 167–172 DOI: 10.1007/BF00278343.

521 (23) Furey, T. S.; Haussler, D. Integration of the cytogenetic map with the draft human

522 genome sequence. Hum. Mol. Genet. 2003, 12 (9), 1037–1044 DOI:

523 10.1093/hmg/ddg113.

524 (24) Strukov, Y. G.; Belmont, a S. Mitotic chromosome structure: reproducibility of

525 folding and symmetry between sister chromatids. Biophys. J. 2009, 96 (4), 1617–

526 1628 DOI: 10.1016/j.bpj.2008.10.051.

527 (25) El-Alfy, M.; Leblond, C. P. An electron microscopic study of mitosis in mouse

528 duodenal crypt cells confirms that the prophasic condensation of chromatin

529 begins during the DNA-synthesizing (S) stage of the cycle. Am. J. Anat. 1989, 186

530 (1), 69–84 DOI: 10.1002/aja.1001860106.

531 (26) Sarkar, A.; Eroglu, S.; Poirier, M. G.; Gupta, P.; Nemani, A.; Marko, J. F.

532 Dynamics of chromosome compaction during mitosis. Exp. Cell Res. 2002, 277

533 (1), 48–56 DOI: 10.1006/excr.2002.5507.

534 (27) Kireeva, N.; Lakonishok, M.; Kireev, I.; Hirano, T.; Belmont, A. S. Visualization of

535 early chromosome condensation: a hierarchical folding, axial glue model of

536 chromosome structure. J. Cell Biol. 2004, 166 (6), 775–785 DOI:

537 10.1083/jcb.200406049. 22 538 (28) Maeshima, K.; Eltsov, M.; Laemmli, U. K. Chromosome structure: improved

539 immunolabeling for electron microscopy. Chromosoma 2005, 114 (5), 365–375

540 DOI: 10.1007/s00412-005-0023-7.

541 (29) Maeshima, K.; Laemmli, U. K. A Two-Step Scaffolding Model for Mitotic

542 Chromosome Assembly. Dev. Cell 2003, 4 (4), 467–480 DOI: 10.1016/S1534-

543 5807(03)00092-3.

544 (30) Marsden, M.; Laemmli, U. Metaphase chromosome structure: Evidence for a

545 radial loop model. Cell 1979, 17 (4), 849–858 DOI: 10.1016/0092-8674(79)90325-

546 8.

547 (31) Liang, Z.; Zickler, D.; Prentiss, M.; Chang, F. S.; Witz, G.; Maeshima, K.;

548 Kleckner, N. Chromosomes Progress to Metaphase in Multiple Discrete Steps via

549 Global Compaction/Expansion Cycles. Cell 2015, 161 (5), 1124–1137 DOI:

550 10.1016/j.cell.2015.04.030.

551 (32) Kleckner, N. Chiasma formation: chromatin/axis interplay and the role(s) of the

552 synaptonemal complex. Chromosoma 2006, 115 (3), 175–194 DOI:

553 10.1007/s00412-006-0055-7.

554 (33) Poirier, M. G.; Marko, J. F. Mitotic chromosomes are chromatin networks without

555 a mechanically contiguous protein scaffold. Proc. Natl. Acad. Sci. U. S. A. 2002,

556 99 (24), 15393–15397 DOI: 10.1073/pnas.232442599.

557 (34) Oliveira, R. A.; Heidmann, S.; Sunkel, C. E. Condensin I binds chromatin early in

558 prophase and displays a highly dynamic association with Drosophila mitotic

559 chromosomes. Chromosoma 2007, 116 (3), 259–274 DOI: 10.1007/s00412-007-

560 0097-5.

561 (35) Schwarzacher, H. G. Chromosomes: in Mitosis and Interphase; Springer Berlin

562 Heidelberg: Berlin, Heidelberg, 1976; pp 8–15.

563 (36) Savvidou, E.; Cobbe, N.; Steffensen, S.; Cotterill, S.; Heck, M. M. S. Drosophila 23 564 CAP-D2 is required for condensin complex stability and resolution of sister

565 chromatids. J. Cell Sci. 2005, 118 (11), 2529–2543 DOI: 10.1242/jcs.02392.

566 (37) Steffensen, S.; Coelho, P. A.; Cobbe, N.; Vass, S.; Costa, M.; Hassan, B.;

567 Prokopenko, S. N.; Bellen, H.; Heck, M. M. S.; Sunkel, C. E. A role for Drosophila

568 SMC4 in the resolution of sister chromatids in mitosis. Curr. Biol. 2001, 11 (5),

569 295–307 DOI: 10.1016/S0960-9822(01)00096-3.

570 (38) Hagstrom, K. A.; Holmes, V. F.; Cozzarelli, N. R.; Meyer, B. J. C. elegans

571 condensin promotes mitotic chromosome architecture, centromere organization,

572 and sister chromatid segregation during mitosis and meiosis. Genes Dev. 2002,

573 16 (6), 729–742 DOI: 10.1101/gad.968302.

574 (39) Ono, T.; Yamashita, D.; Hirano, T. Condensin II initiates sister chromatid

575 resolution during S phase. J. Cell Biol. 2013, 200 (4), 429–441 DOI:

576 10.1083/jcb.201208008.

577 (40) Sikorav, J. L.; Jannink, G. Kinetics of chromosome condensation in the presence

578 of topoisomerases: a phantom chain model. Biophys. J. 1994, 66 (3), 827–837

579 DOI: 10.1016/S0006-3495(94)80859-8.

580 (41) DiNardo, S.; Voelkel, K.; Sternglanz, R. DNA topoisomerase II mutant of

581 Saccharomyces cerevisiae: topoisomerase II is required for segregation of

582 daughter molecules at the termination of DNA replication. Proc. Natl. Acad. Sci.

583 1984, 81 (9 ), 2616–2620.

584 (42) Newport, J.; Spann, T. Disassembly of the nucleus in mitotic extracts: Membrane

585 vesicularization, lamin disassembly, and chromosome condensation are

586 independent processes. Cell 1987, 48 (2), 219–230 DOI: 10.1016/0092-

587 8674(87)90425-9.

588 (43) Shamu, C. E.; Murray, A. W. Sister chromatid separation in frog egg extracts

589 requires DNA topoisomerase II activity during anaphase. J. Cell Biol. 1992, 117 24 590 (5), 921–934 DOI: 10.1083/jcb.117.5.921.

591 (44) Kueng, S.; Hegemann, B.; Peters, B. H.; Lipp, J. J.; Schleiffer, A.; Mechtler, K.;

592 Peters, J.-M. Wapl controls the dynamic association of cohesin with chromatin.

593 Cell 2006, 127 (5), 955–967 DOI: 10.1016/j.cell.2006.09.040.

594 (45) Gandhi, R.; Gillespie, P. J.; Hirano, T. Human Wapl Is a Cohesin-Binding Protein

595 that Promotes Sister-Chromatid Resolution in Mitotic Prophase. Curr. Biol. 2006,

596 16 (24), 2406–2417 DOI: http://dx.doi.org/10.1016/j.cub.2006.10.061.

597 (46) Nasmyth, K.; Haering, C. H. Cohesin: Its Roles and Mechanisms. Annu. Rev.

598 Genet. 2009, 43 (1), 525–558 DOI: 10.1146/annurev-genet-102108-134233.

599 (47) Birshtein, T. M.; Borisov, O. V.; Zhulina, Y. B.; Khokhlov, A. R.; Yurasova, T. A.

600 Conformations of comb-like macromolecules. Polym. Sci. U.S.S.R. 1987, 29 (6),

601 1293–1300 DOI: 10.1016/0032-3950(87)90374-1.

602 (48) Ball, R. C.; Marko, J. F.; Milner, S. T.; Witten, T. A. Polymers grafted to a convex

603 surface. Macromolecules 1991, 24 (3), 693–703 DOI: 10.1021/ma00003a011.

604 (49) Li, H.; Witten, T. A. Polymers Grafted to Convex Surfaces - a Variational

605 Approach. Macromolecules 1994, 27 (2), 449–457 DOI: 10.1021/ma00080a019.

606 (50) Lawrimore, J.; Vasquez, P. A.; Falvo, M. R.; Taylor, R. M.; Vicci, L.; Yeh, E.;

607 Forest, M. G.; Bloom, K. DNA loops generate intracentromere tension in mitosis.

608 J. Cell Biol. 2015, 210 (4), 553–564 DOI: 10.1083/jcb.201502046.

609 (51) Lawrimore, J.; Aicher, J. K.; Hahn, P.; Fulp, A.; Kompa, B.; Vicci, L.; Falvo, M.;

610 Taylor, R. M.; Bloom, K. ChromoShake: a chromosome dynamics simulator

611 reveals that chromatin loops stiffen centromeric chromatin. Mol. Biol. Cell 2016,

612 27 (1), 153–166 DOI: 10.1091/mbc.E15-08-0575.

613 (52) Zhang, Y.; Heermann, D. W. Loops Determine the Mechanical Properties of

614 Mitotic Chromosomes. PLoS One 2011, 6 (12), e29225 DOI:

615 10.1371/journal.pone.0029225. 25 616 (53) Cheng, T. M. K.; Heeger, S.; Chaleil, R. A. G.; Matthews, N.; Stewart, A.; Wright,

617 J.; Lim, C.; Bates, P. A.; Uhlmann, F. A simple biophysical model emulates

618 budding yeast chromosome condensation. Elife 2015, 4 DOI:

619 10.7554/eLife.05565.

620 (54) Wang, X.; Le, T. B. K.; Lajoie, B. R.; Dekker, J.; Laub, M. T.; Rudner, D. Z.

621 Condensin promotes the juxtaposition of DNA flanking its loading site in Bacillus

622 subtilis. Genes Dev. 2015, 29 (15 ), 1661–1675 DOI: 10.1101/gad.265876.115.

623 (55) Wilhelm, L.; Bürmann, F.; Minnen, A.; Shin, H.-C.; Toseland, C. P.; Oh, B.-H.;

624 Gruber, S. SMC condensin entraps chromosomal DNA by an ATP hydrolysis

625 dependent loading mechanism in Bacillus subtilis. Elife 2015, 4 DOI:

626 10.7554/eLife.06659.

627 (56) Mora-Bermúdez, F.; Gerlich, D.; Ellenberg, J. Maximal chromosome compaction

628 occurs by axial shortening in anaphase and depends on Aurora kinase. Nat. Cell

629 Biol. 2007, 9 (7), 822–831 DOI: 10.1038/ncb1606.

630 (57) Leblond, C. P.; El-Alfy, M. The eleven stages of the , with emphasis on

631 the changes in chromosomes and nucleoli during interphase and mitosis. Anat.

632 Rec. 1998, 252 (3), 426–443 DOI: 10.1002/(SICI)1097-

633 0185(199811)252:3<426::AID-AR11>3.0.CO;2-3.

634 (58) Hirota, T.; Gerlich, D.; Koch, B.; Ellenberg, J.; Peters, J.-M. Distinct functions of

635 condensin I and II in mitotic chromosome assembly. J. Cell Sci. 2004, 117 (Pt 26),

636 6435–6445 DOI: 10.1242/jcs.01604.

637 (59) Dietzel, S.; Belmont, A. S. Reproducible but dynamic positioning of DNA in

638 chromosomes during mitosis. Nat Cell Biol 2001, 3 (8), 767–770 DOI:

639 10.1038/35087089.

640 (60) Hirano, T.; Mitchison, T. J. Cell cycle control of higher-order chromatin assembly

641 around naked DNA in vitro. J. Cell Biol. 1991, 115 (6), 1479–1489 DOI: 26 642 10.1083/jcb.115.6.1479.

643 (61) Gillespie, D. T. Exact stochastic simulation of coupled chemical reactions. J.

644 Phys. Chem. 1977, 81 (25), 2340–2361 DOI: 10.1021/j100540a008.

645 (62) Eastman, P.; Pande, V. S. OpenMM: A Hardware Independent Framework for

646 Molecular Simulations. Comput. Sci. Eng. 2010, 12 (4), 34–39 DOI:

647 10.1109/MCSE.2010.27.

648 (63) Eastman, P.; Friedrichs, M. S.; Chodera, J. D.; Radmer, R. J.; Bruns, C. M.; Ku, J.

649 P.; Beauchamp, K. a; Lane, T. J.; Wang, L.-P.; Shukla, D.; et al. OpenMM 4: A

650 Reusable, Extensible, Hardware Independent Library for High Performance

651 Molecular Simulation. J. Chem. Theory Comput. 2013, 9 (1), 461–469 DOI:

652 10.1021/ct300857j.

653 (64) Le, T. B. K.; Imakaev, M. V; Mirny, L. A.; Laub, M. T. High-Resolution Mapping of

654 the Spatial Organization of a Bacterial Chromosome. Science (80-. ). 2013, 342

655 (6159), 731–734 DOI: 10.1126/science.1242059.

656

27 657 FIGURES

658 Figure 1. Model of loop extrusion by condensins. Top row, the update rules used in

659 simulations: (A) a condensin extrudes a loop by moving the two ends along the chromosome in

660 the opposite directions, (B) collision of condensins bound to chromosomes blocks loop

661 extrusion on the collided sides, (C) a condensin spontaneously dissociates and the loop

662 disassembles; (D) a condensin associates at randomly chosen site and starts extruding a loop.

663 Bottom row, (E) we study how combined action of many loop extruding condensins changes the

664 conformation of a long chromosome using polymer simulations.

665

666 Figure 2. Loop-extruding condensins convert globular interphase chromosomes into

667 elongated structures. (A) Dynamics of loop array formation. Left, the fraction of the

668 chromosome extruded into loops (loop coverage, gray) and the mean loop size (black). The

669 time is measured in the units of condensin residence time. Data for each curve was obtained

670 using 10 simulation replicas; the thickness of each line shows the difference between the 10th

671 and 90th percentile at each time point. Right, a phase diagram showing loop coverage vs mean

672 loop length. The curve is averaged over the simulation replicas and colored by log-time. Two

673 stages of dynamics are separated by the two vertical dashed lines corresponding to 95%

674 coverage and 95% of the steady state mean loop length. The distinction between the two stages

675 of compaction is especially easy to see on the phase curve. (B) Dynamics of chromosome

676 morphology. Left: chromosome length (blue) and the mean chromosome diameter (red) shown

677 as a function of time; the thickness of each line shows the difference between the 10th and 90th

678 percentile at each time point. Right, a phase diagram showing chromosome length vs diameter;

679 the curve is averaged over the simulation replicas and colored by log-time. Note that most of

680 chromosome linear compaction is achieved during the first phase, while the second is

681 characterized by widening of the chromosome and further three-fold shortening. (C) Snapshots

682 of polymer simulations. Top row: the conformations of chromosomes at the beginning of

683 simulations (left), slightly before the transition between the loop formation and maturation stages

684 (middle), and in the steady state (right). Color (rainbow: from red to blue) shows positions of 28 685 individual monomers along the chain and demonstrates that loci in compacted chromosomes

686 are arranged linearly according to their genomic location. Bottom row: same as above, with

687 chromosomes shown in semitransparent gray and condensins as red spheres.

688

689 Figure 2 - figure supplement 1. The geometrical parameters of compacted chromosomes

690 in simulations with altered parameters. (A) Simulations with altered condensin abundance:

691 “0.2x condensin”, “0.5x condensin”, “2x condensins” and “4x condensins” - simulations with 200,

692 500, 2000 and 4000 condensins, correspondingly. (B) Simulations with altered parameters of

693 the chromatin fiber and condensins: “Stiffer fiber” – simulations with 2x bending energy of the

694 fiber, “softer fiber” – simulations with 0.5x bending energy; “wide condensins” – simulations with

695 increased spacing between condensins at loop bases; “10nm fiber” - simulations with reduced

696 linear DNA of the fiber; “30nm fiber” – simulations with increased linear DNA density of the fiber

697 and increased fiber thickness. The vertical gray lines show the experimental observations on the

698 linear density of DNA along the human prophase chromosome I (6.25 kb/nm) and chromosome

699 XXII (5.0 kb/nm) 20. The horizontal gray lines show the experimental data on the diameter of

700 prophase chromosomes: 400 nm 27, 600 nm 26 and 679 nm 25.

701

702 Figure 3. Segregation and disentanglement of sister chromatids. (A) The median

703 separation between condensin-rich axes of two sister chromosomes (blue) and the mean

704 chromosome diameter (red) as a function of time, measured in condensin residence times, as

705 above. The blue band shows the range between 25th and 75th percentile across 10 simulation

706 replicas at each time point; the light blue shows 10%--90% range. (B) Entanglement of sister

707 chromatids measured by the linking number of the condensin cores of sister chromatids; the

708 blue bands show variability between simulation replicas, same as above. (C) Conformations of

709 two sister chromatids in the polymer simulations (shown in red and blue) illustrate the observed

710 segregation and disentanglement.

711

29 712 Figure 3 - figure supplement 1. Segregation and disentanglement of sister chromatids in

713 simulations with reduced Topo II concentration. (A) The median separation between

714 condensin-rich axes of two sister chromosomes (solid lines) and the mean chromosome

715 diameter (red) as a function of time, measured in condensin residence times, as above.

716 Depletion of topo II was simulated by a simultaneous increase of the energy barrier of fiber

717 overlap (shown by the line color) and a 3x increase in of the radius of repulsion. (B)

718 Entanglement of sister chromatids in simulations of topo II depletion, measured by the linking

719 number of the condensin cores of sister chromatids.

720

721 Figure 4. Steric repulsion between chromatin loops shapes the compacted chromosomes

722 and leads to segregation of sister chromatids. (A) Left, conformation of a single

723 chromosomes with chromatin fiber shown in semitransparent gray and condensins shown in

724 red; two loops are highlighted with blue. Right: the diagram showing that loop repulsion

725 straightens chromosomal cores and extends individual loops radially. (B) Left, cross section of

726 two parallel loop brushes placed at a short distance from each other. Right, the diagram shows

727 that steric interactions between loops of sister chromatids lead to chromatid repulsion and

728 segregation.

729

730 VIDEO CAPTIONS

731

732 Video 1: An illustration of the model of loop extrusion by condensins. When a condensin

733 (red sphere) binds to chromatin (grey fiber), it binds to two adjacent loci and then slides the two

734 contact points in the opposite directions, effectively extruding a loop (highlighted in orange). The

735 diagram on the bottom shows the position of the two loci brought together by the condensin

736 over time (red arc). Condensins occasionally unbind and then rebind to a randomly chosen site,

737 starting a new loop. When two condensins (loops shown in orange and purple) meet on the

738 chromosome, they stop extrusion on the collided sides. Condensins can also bind within already

739 extruded loops. If the growth of the outer loop is blocked by other loops or barriers (not shown), 30 740 the inner condensin can re-extrude it, forming a single loop reinforced by multiple condensins

741 (the number above the arc). Such a loop is stabilized against unbinding of individual condensins

742 and disassembles only upon unbinding of the last condensin. Available at

743 https://www.youtube.com/watch?v=xikxZ0yF1yU&list=PLnV-

744 JMzdy0ZMOvyK6Z1bzU7C2Jt8Brbit&index=1

745

746 Video 2: A representative polymer simulation of compaction of 30 Mb chromosome by

747 loop extruding condensins in presence of strand-passing topo II. Top left, fiber color

748 corresponds to the position of each monomer along the genome. Top right, fiber shown in

749 semitransparent gray, condensins are shown as red spheres. Bottom, the diagram of the

750 crosslinks (red arcs) formed by condensins. The numbers above the arcs show the number of

751 condensins stacked at the bases of reinforced loops. We sped up the second half of the video

752 20 times to illustrate the slow phase of loop maturation and the long-term stability of the

753 compacted state. Available at https://www.youtube.com/watch?v=_Vc7__xfnfc&list=PLnV-

754 JMzdy0ZMOvyK6Z1bzU7C2Jt8Brbit&index=2

755

756 Video 3: A polymer simulation of segregation and disentanglement of two sister

757 chromatids by loop extruding condensins in presence of strand-passing topo II. The two

758 sister chromatids (30Mb each) are shown in red and blue. Initially, the two chromatids were

759 compacted and twisted around each other. Second half of the video is sped up 20x. Available at

760 https://www.youtube.com/watch?v=stZR5s9n32s&index=3&list=PLnV-

761 JMzdy0ZMOvyK6Z1bzU7C2Jt8Brbit

762

763

764 Video 4: Sixteen simulation replicas of chromatid segregation and disentanglement.

765 Available at https://www.youtube.com/watch?v=UXXLnGGit-8&index=4&list=PLnV-

766 JMzdy0ZMOvyK6Z1bzU7C2Jt8Brbit

767

31 768 Video 5: Polymer simulation of segregation and disentanglement of two sister

769 chromatids with altered model parameters. Top row: left, simulations with a more flexible

770 fiber (0.5x bending energy); right – a stiffer fiber (2x bending energy); bottom row, left -

771 simulations with wider condensins at loop bases; right – simulations with disabled excluded

772 volume interactions. Available at

773 https://www.youtube.com/watch?v=YhgSHppJXVI&index=5&list=PLnV-

774 JMzdy0ZMOvyK6Z1bzU7C2Jt8Brbit

775

776

777 Video 6: Polymer simulation of segregation and disentanglement of two sister

778 chromatids with modified condensin abundance. Top row, left - 2000 condensins per

779 chromatid, middle - 500 condensins per chromatid; right - 200 condensins; bottom row, left -

780 4000 condensins, left - 100 condensins; right, 50 condensins. Available at

781 https://www.youtube.com/watch?v=oq9TS5OXO1Q&index=6&list=PLnV-

782 JMzdy0ZMOvyK6Z1bzU7C2Jt8Brbit

783

784

785 Video 7: A polymer simulation of segregation and disentanglement of two sister

786 chromatids upon topo II depletion. Depletion of topo II was simulated by a simultaneous

787 increase of the energy barrier of fiber overlap and the increased radius of repulsion, which

788 decreased, but not fully prevented chain crossings. Available at

789 https://www.youtube.com/watch?v=YO3Qx8liK3s&list=PLnV-

790 JMzdy0ZMOvyK6Z1bzU7C2Jt8Brbit&index=7

791

792 Video 8: A polymer simulation of segregation and disentanglement of sister chromatids

793 in the presence of cohesins connecting the chromatids. We simulated cohesins as

794 crosslinks between the chromosomes (shown with the blue lines on the bottom arc diagram)

795 that can be pushed along each of the chromatids by loop-extruding condensins. Available at

32 796 https://www.youtube.com/watch?v=vmwNbz41pqM&list=PLnV-

797 JMzdy0ZMOvyK6Z1bzU7C2Jt8Brbit&index=8

798

33