Entropy-driven spatial organization of highly confined polymers: … · 2006. 9. 22. · presence...

6
Entropy-driven spatial organization of highly confined polymers: Lessons for the bacterial chromosome Suckjoon Jun and Bela Mulder Stichting voor Fundamenteel Onderzoek der Materie (FOM) Institute for Atomic and Molecular Physics (AMOLF), Kruislaan 407, 1098 SJ, Amsterdam, The Netherlands Communicated by Nancy Kleckner, Harvard University, Cambridge, MA, June 29, 2006 (received for review April 25, 2006) Despite recent progress in visualization experiments, the mecha- nism underlying chromosome segregation in bacteria still remains elusive. Here we address a basic physical issue associated with bacterial chromosome segregation, namely the spatial organiza- tion of highly confined, self-avoiding polymers (of nontrivial to- pology) in a rod-shaped cell-like geometry. Through computer simulations, we present evidence that, under strong confinement conditions, topologically distinct domains of a polymer complex effectively repel each other to maximize their conformational entropy, suggesting that duplicated circular chromosomes could partition spontaneously. This mechanism not only is able to ac- count for the spatial separation per se but also captures the major features of the spatiotemporal organization of the duplicating chromosomes observed in Escherichia coli and Caulobacter crescentus. bacterial chromosome segregation Caulobacter crescentus Escherichia coli polymer physics B acteria are arguably the simplest living organisms that are able to reproduce independently. The key step in cellular reproduction consists of the duplication of the genetic material and its subsequent distribution over the offspring. Successful as they are from an evolutionary perspective, bacteria are indeed considered to be highly efficient and accurate in these processes. What then is the mechanism underlying bacterial chromosome segregation? Perhaps the most influential model so far was proposed by Jacob et al. in 1963 (1) in their seminal paper on the replicon model of Escherichia coli; if replicating chromosomes are attached to the elongating cell-wall membrane, they can be segregated passively by insertion of membrane material between the attachment points. Although the hypothesis by Jacob et al. (1) is intuitive and elegant, it has been undermined by a series of recent visualization experiments. In Bacillus subtilis (2) and Caulobacter crescentus (3), for example, the measured rate of movement of the replication origin ranges from 0.1 to 0.3 mmin, almost 10 times larger than the cell elongation rate. In E. coli, on the other hand, the movement rate of DNA segments is still a matter of debate [see, for example, Elmore et al. (4)]. Nevertheless, all three species, including E. coli, show striking dynamics and directed longitudinal movements of chromosome loci during replication; although details vary from organism to organism, typically the duplicated replication origins (ori) move toward opposite poles in the cell, and the terminus (ter) moves toward the cell center (3, 5–7). A number of alternative models have been proposed recently to explain the driving force for DNA segregation in bacteria [for recent reviews, see Woldringh (8), Errington et al. (9), and Gitai et al. (10) and references therein]. A consensus view that seems to emerge from this more recent literature, at least in case of C. crescentus, favors the existence of putative ‘‘eukaryotic-like’’ mechanisms (11–13), which would actively pushpull the dupli- cating bacterial chromosomes apart. However, Marko and coworkers (14, 15) and, more recently and more specifically, Kleckner and coworkers (7, 16) have drawn attention to the possible role of forces with a purely physical origin related to the mechanical state of the DNA. A potential source for these forces derives from the fact that DNA as a polymer can both store and release entropy, allowing it to perform mechanical work. Despite this fundamental property, most of the current models for DNA segregation do not take into account the entropy associated with the chromosomal confor- mations. In fact, we are not aware of any quantitative studies aimed at elucidating these effects. Moreover, the widespread conviction that an active segregation mechanism is in fact a necessity seems to derive in part from the implicit assumption that the duplicated DNA would by itself remain completely intermingled. This assumption, which to our knowledge has never been stated explicitly, still remains to be tested. In the present work, we therefore address the basic physics behind these biologically relevant questions and explore several aspects of the spatial organization and segregation of strongly confined entropic polymers using computer simulations as well as scaling arguments, whenever possible. Because the duplicating bacterial chromosome can be regarded as an example of a closed polymer of slowly evolving topology, we are then naturally led to the question of what insights polymer physics can provide on the bacterial DNA segregation process. As we shall show below, the combination of confinement and topological constraints that occurs in these systems produces spatially structured equilibrium states of the polymers considered, which bear a striking resem- blance to the ones actually found for the bacterial chromosome. Moreover, we also demonstrate how compaction of the bacterial chromosome and conformational entropy alone could direct and facilitate the movement of newly replicated daughter strands of DNA. Results and Discussion Role of Entropy in Polymer Segregation in Bulk and in Confinement. When two polymeric coils consisting of N monomers, each in dilute solution (in a good solvent), are brought together within a distance smaller than their natural size, given by the radius of gyration R g N ( 35 is the Flory constant) (17), both coils lose some of their conformational entropy because of the excluded-volume interactions and the chain connectivity and, thus, will resist overlap. This effective entropic repulsion be- tween two chains was first characterized by Flory and Krigbaum in 1950 (18), who by using a self-consistent field theory proposed that, when the two chains completely overlap, the free energy cost (F) of overlap scales as F Flory N 1/5 k B T k B T, thus predicting that long coils should behave effectively as mutually impenetrable hard spheres. Much later, however, based on the results of des Cloizeaux (19) and Daoud et al. (20), Grosberg et al. (21) showed with scaling arguments that in fact F k B T and thus is both weak and independent of the chain length N in the Conflict of interest statement: No conflicts declared. Freely available online through the PNAS open access option. To whom correspondence should be addressed. E-mail: [email protected]. 1kBT 4 10 21 J 10 24 kcal is a typical scale of thermal energy at room temperature. © 2006 by The National Academy of Sciences of the USA 12388 –12393 PNAS August 15, 2006 vol. 103 no. 33 www.pnas.orgcgidoi10.1073pnas.0605305103 Downloaded by guest on March 5, 2021

Transcript of Entropy-driven spatial organization of highly confined polymers: … · 2006. 9. 22. · presence...

Page 1: Entropy-driven spatial organization of highly confined polymers: … · 2006. 9. 22. · presence of strong confinement. Simulations of two ring poly-mers, in the same setting as

Entropy-driven spatial organization of highly confinedpolymers: Lessons for the bacterial chromosomeSuckjoon Jun† and Bela Mulder

Stichting voor Fundamenteel Onderzoek der Materie (FOM) Institute for Atomic and Molecular Physics (AMOLF), Kruislaan 407,1098 SJ, Amsterdam, The Netherlands

Communicated by Nancy Kleckner, Harvard University, Cambridge, MA, June 29, 2006 (received for review April 25, 2006)

Despite recent progress in visualization experiments, the mecha-nism underlying chromosome segregation in bacteria still remainselusive. Here we address a basic physical issue associated withbacterial chromosome segregation, namely the spatial organiza-tion of highly confined, self-avoiding polymers (of nontrivial to-pology) in a rod-shaped cell-like geometry. Through computersimulations, we present evidence that, under strong confinementconditions, topologically distinct domains of a polymer complexeffectively repel each other to maximize their conformationalentropy, suggesting that duplicated circular chromosomes couldpartition spontaneously. This mechanism not only is able to ac-count for the spatial separation per se but also captures the majorfeatures of the spatiotemporal organization of the duplicatingchromosomes observed in Escherichia coli and Caulobactercrescentus.

bacterial chromosome segregation � Caulobacter crescentus �Escherichia coli � polymer physics

Bacteria are arguably the simplest living organisms that areable to reproduce independently. The key step in cellular

reproduction consists of the duplication of the genetic materialand its subsequent distribution over the offspring. Successful asthey are from an evolutionary perspective, bacteria are indeedconsidered to be highly efficient and accurate in these processes.

What then is the mechanism underlying bacterial chromosomesegregation? Perhaps the most influential model so far wasproposed by Jacob et al. in 1963 (1) in their seminal paper on thereplicon model of Escherichia coli; if replicating chromosomesare attached to the elongating cell-wall membrane, they can besegregated passively by insertion of membrane material betweenthe attachment points. Although the hypothesis by Jacob et al.(1) is intuitive and elegant, it has been undermined by a seriesof recent visualization experiments. In Bacillus subtilis (2) andCaulobacter crescentus (3), for example, the measured rate ofmovement of the replication origin ranges from 0.1 to 0.3�m�min, almost 10 times larger than the cell elongation rate. InE. coli, on the other hand, the movement rate of DNA segmentsis still a matter of debate [see, for example, Elmore et al. (4)].Nevertheless, all three species, including E. coli, show strikingdynamics and directed longitudinal movements of chromosomeloci during replication; although details vary from organism toorganism, typically the duplicated replication origins (ori) movetoward opposite poles in the cell, and the terminus (ter) movestoward the cell center (3, 5–7).

A number of alternative models have been proposed recentlyto explain the driving force for DNA segregation in bacteria [forrecent reviews, see Woldringh (8), Errington et al. (9), and Gitaiet al. (10) and references therein]. A consensus view that seemsto emerge from this more recent literature, at least in case of C.crescentus, favors the existence of putative ‘‘eukaryotic-like’’mechanisms (11–13), which would actively push�pull the dupli-cating bacterial chromosomes apart.

However, Marko and coworkers (14, 15) and, more recentlyand more specifically, Kleckner and coworkers (7, 16) havedrawn attention to the possible role of forces with a purely

physical origin related to the mechanical state of the DNA. Apotential source for these forces derives from the fact that DNAas a polymer can both store and release entropy, allowing it toperform mechanical work. Despite this fundamental property,most of the current models for DNA segregation do not take intoaccount the entropy associated with the chromosomal confor-mations. In fact, we are not aware of any quantitative studiesaimed at elucidating these effects. Moreover, the widespreadconviction that an active segregation mechanism is in fact anecessity seems to derive in part from the implicit assumptionthat the duplicated DNA would by itself remain completelyintermingled. This assumption, which to our knowledge hasnever been stated explicitly, still remains to be tested.

In the present work, we therefore address the basic physicsbehind these biologically relevant questions and explore severalaspects of the spatial organization and segregation of stronglyconfined entropic polymers using computer simulations as wellas scaling arguments, whenever possible. Because the duplicatingbacterial chromosome can be regarded as an example of a closedpolymer of slowly evolving topology, we are then naturally led tothe question of what insights polymer physics can provide on thebacterial DNA segregation process. As we shall show below, thecombination of confinement and topological constraints thatoccurs in these systems produces spatially structured equilibriumstates of the polymers considered, which bear a striking resem-blance to the ones actually found for the bacterial chromosome.Moreover, we also demonstrate how compaction of the bacterialchromosome and conformational entropy alone could direct andfacilitate the movement of newly replicated daughter strandsof DNA.

Results and DiscussionRole of Entropy in Polymer Segregation in Bulk and in Confinement.When two polymeric coils consisting of N monomers, each indilute solution (in a good solvent), are brought together withina distance smaller than their natural size, given by the radius ofgyration Rg � N� (� � 3�5 is the Flory constant) (17), both coilslose some of their conformational entropy because of theexcluded-volume interactions and the chain connectivity and,thus, will resist overlap. This effective entropic repulsion be-tween two chains was first characterized by Flory and Krigbaumin 1950 (18), who by using a self-consistent field theory proposedthat, when the two chains completely overlap, the free energycost (F) of overlap scales as FFlory � N1/5kBT �� kBT,‡ thuspredicting that long coils should behave effectively as mutuallyimpenetrable hard spheres. Much later, however, based on theresults of des Cloizeaux (19) and Daoud et al. (20), Grosberg etal. (21) showed with scaling arguments that in fact F � kBT andthus is both weak and independent of the chain length N in the

Conflict of interest statement: No conflicts declared.

Freely available online through the PNAS open access option.

†To whom correspondence should be addressed. E-mail: [email protected].

‡1kBT � 4 � 10�21 J � 10�24 kcal is a typical scale of thermal energy at room temperature.

© 2006 by The National Academy of Sciences of the USA

12388–12393 � PNAS � August 15, 2006 � vol. 103 � no. 33 www.pnas.org�cgi�doi�10.1073�pnas.0605305103

Dow

nloa

ded

by g

uest

on

Mar

ch 5

, 202

1

Page 2: Entropy-driven spatial organization of highly confined polymers: … · 2006. 9. 22. · presence of strong confinement. Simulations of two ring poly-mers, in the same setting as

scaling regime, which was later confirmed by renormalizationgroup calculations (22) and numerical simulations (refs. 23 and24; for a review, see ref. 25). Indeed, DNA molecules in bulk caneasily intermingle.

In the presence of confinement, however, the effective repul-sion can increase dramatically. To see this effect, consider twolinear chains with excluded-volume interactions, which are ini-tially intermingled and confined in an infinitely long cylinder ofdiameter D �� Rg (Fig. 1). The free energy cost of interminglingin this case can be estimated using de Gennes’ ‘‘blob’’ model (17).Because the total number of blobs per chain is of order D�(1/�)N,the overlapping free energy scales as

Fblob � D�1� NkBT , [1]

where we assumed that the individual blob–blob overlap cost isof order kBT based on the abovementioned results by Grosberget al. (21). The repulsion between two chains in confinement isthus predicted to be very strong, being proportional to the chainlength N, in stark contrast to the almost negligibly weak one inbulk. Because of this entropic driving force, the two chains slideaway from each other at a characteristic rate v* � kBT�DN overa distance comparable with the size of an unperturbed chainl* � D1�(1/�)N, giving the timescale of segregation as �seg �l*�v* � N2D2�(1/�). Note that this �seg is much smaller (faster)than the typical timescale, �rept � N3D2�(2/�), for a single chainin the same cylindrical confinement to diffuse over a distance ofits own length by ‘‘reptation-like’’ motion.

Demixing of Chains in a Closed Geometry of Confinement and theEffect of Chain Topology. Although the above analysis in an opencylinder clearly demonstrates the effect of lateral confinementon chain segregation, the biologically more relevant geometry isthe one in which also the longitudinal direction is confined as ina generically rod-shaped bacterium. We thus performed MonteCarlo simulations to find the typical equilibrium configuration oftwo linear chains in a closed cylinder of length L capped by twohemispheres of diameter D, where the aspect ratio (L � D)�Dof 6 and the volume fraction of the chain �10% are chosen tobe comparable with the situation in slowly growing E. coli nearthe onset of cell division (26).§ The results depicted in Fig. 2Ashow that the majority of the chains in our limited sampling ofinitial conditions end up being fully demixed (for further detailson the Monte Carlo simulations, see Supporting Text, which ispublished as supporting information on the PNAS web site).

The bacterial chromosome, however, is not a linear polymerbut a circular one. In bulk solutions, it is known that there is anadditional topological repulsion between ring polymers due tothe nonconcatenation condition (27). On the other hand, thecircular topology might only have a secondary effect in thepresence of strong confinement. Simulations of two ring poly-

mers, in the same setting as the linear ones, indicate that, ifanything, two ring polymers demix more readily than linear onesdo in confinement (Fig. 2B).

Finally, the bacterial chromosome is also supercoiled, due tothe undertwisting of the DNA-helix. This supercoiling causes thestructure of the chromosome to resemble a random branchingnetwork composed of plectonemes with a length of �300 nm. Wetherefore also considered the mixing behavior of pairs ofbranched polymers under confinement. The results shown in Fig.2C suggest that the nontrivial topology of the branched polymersincreases the tendency to demix even further.

Although clearly not in any way a systematic study of theeffects discussed above, which at the present state-of-the-artwould require a major commitment in computational resources,we contend that the trends clearly reveal that both confinementand chain topology of the bacterial chromosome impact thesegregation behavior.

Spatial Organization of Polymers Under Confinement. As mentionedabove, during the cell cycle, a circular bacterial chromosomeundergoes a series of topological and geometrical changes (Fig.3A). Given the observations on the directed movements ofduplicating bacterial chromosome, it is of interest to understandthe typical conformation of a highly confined chain in a rod-shaped box, where the chain topology and volume fraction mimicthat of the bacterial chromosome in different stages of DNAreplication.

The results from our simulations, summarized in Fig. 3 B andC, suggest how conformational entropy might lead to an evolu-tion of spatiotemporal organization of duplicating circular chainin an elongating rod-shaped cell: In the early stage of replication,the replication ‘‘bubble’’ is small (asymmetric figure-�) andextruded from the mother strand that has not yet been repli-cated. At this stage the duplicated ori and ter are localized nearthe opposite poles. The reason for this ori localization is adepletion effect; when an object whose size is comparable withor larger than the thickness of a polymer is inserted into a volumeoccupied by a long polymer chain, the chain ‘‘expels’’ this objectto gain more entropy. Thus, at the beginning of the cell cycle, thereplication bubble formed by the two daughter strands does notmix with the long mother strand. Moreover, as can be seen in Fig.3B, bipolarity is induced by the geometry of the rod-shaped box,with both the daughter strands tending to occupy a volume at oneof the two poles. Similarly, the midpoint of the mother strand(namely, ter) will therefore be located, on average, near the otherpole. Next, when the circular chromosome is �50% replicated,the mother strand and the two daughter strands are all approx-imately equal in length (symmetric figure-�); i.e., the chain canbe divided into three topologically distinct domains of identical

§The confinement in our simulations is moderately strong, e.g., the radius of gyration Rg ofeach linear chain is �3.6 times larger than the radius of the cylinder D�2.

Fig. 1. Blob approach to segregation of two partly intermingled chains in acylindrical pore. R� � ND1�(1/�) is the longitudinal size of a single chain. The freeenergy of overlap Fblob scales as kBTRoverlap�D.

Fig. 2. Segregation of two chains of various topology in a rod-shaped box oflength Ltot (normalized center-to-center distance between two chains vs.Monte Carlo sweeps). (A) Linear. (B) Ring. (C) Branched. In all three cases, weused a box of the same dimensions and repeated each simulation for 10different initial configurations. The aspect ratio of each graph has beenchosen such that the ‘‘timing’’ of the segregation can be compared.

Jun and Mulder PNAS � August 15, 2006 � vol. 103 � no. 33 � 12389

BIO

PHYS

ICS

Dow

nloa

ded

by g

uest

on

Mar

ch 5

, 202

1

Page 3: Entropy-driven spatial organization of highly confined polymers: … · 2006. 9. 22. · presence of strong confinement. Simulations of two ring poly-mers, in the same setting as

size, which repel each other and are (weakly) compartmentalizedinside the box.

Perhaps the most interesting case is the chain with a figure-eight topology, which occurs in the final stage of replication ofany bacterium having a circular chromosome. Shown in Fig. 3Cis the typical conformation of a figure-eight chain, as well as theaverage positional distribution and fluctuation of each of itsmonomers along the long axis inside the cylinder. The firstimportant feature of the conformation of this figure-eight chainin confinement is its mirror-like, linear ordering within a rod-shaped box. The ter locus (the link of the two rings) is on averagelocated in the center, whereas the two ori loci (the ‘‘apex’’ of eachring) are near the two opposite poles. Connecting these threeloci, there is a precise linear correspondence between the clockpositions of monomers along the chain and their spatial orderinginside the cylinder.

An important observation in Fig. 3C is that the magnitude ofthe positional f luctuations along the longitudinal axis is large andcomparable with the width of the cylinder, i.e., ��zi

21/2 (�zi2 �

�zi2)1/2 � D (where zi is the position of the i-th monomer, andthe brackets � denote the average of what is inside), reminiscentof the blob picture (e.g., Fig. 1).

Simple Model of a Bacterial Chromosome During Replication. So far,we have shown how the interplay between conformationalentropy and chain topology can lead a highly confined polymerto adopt a very specific spatial organization within a rod-shaped geometry. In particular, the spontaneous partitioningof two chains of fixed topology and geometry, as well as thelinear organization of the figure-eight chain, strongly suggestthat conformational entropy per se could be a major drivingforce underlying chromosome segregation. Naturally, we arethen led to ask to what extent these physical principles couldin fact be responsible for segregation of duplicating bacterialchromosomes.

For this goal, we need a computational model of a duplicatingbacterial chromosome, which not only is simple, so that theconsequences of underlying assumptions (be they physicaland�or biological) remain transparent, but that also reflects theessential features of the in vivo situation. We therefore need toaddress the properties of the nucleoid, the distinct intracellulardomain to which the chromosome is confined.

From the segregation point of view, perhaps the most impor-tant aspect of the bacterial nucleoid is its compaction.¶ Forinstance, as mentioned above, supercoiling can change the globalshape of a large circular DNA chain to mimic a more ordered,branched tree-like structure. Moreover, nucleoid-associated andstructural maintenance of chromosome (SMC) proteins in bac-teria (30), as well as molecular crowding (31), could furthercondense the DNA and stabilize its spatial organization. Indeed,in an important work, Sawitzke and Austin (32) demonstratedhow all these factors, DNA condensation, supercoiling, and SMCproteins, are important for chromosome segregation (see alsoref 33); they showed that the failure of proper chromosomesegregation in E. coli due to mutations of the muk genes(structural and functional analogues of SMC family of proteins)can be alleviated by a modest increase in chromosomal super-coiling, supporting the idea that the condensation (and thestructure) of the chromosome is critical for the efficiency of thebacterial chromosome segregation.

Based on the arguments above, we introduce a ‘‘concentric-shell’’ model of bacterial nucleoid (see Fig. 4). Our idea is toconfine the unreplicated mother strands to a rod-shaped ‘‘inner-nucleoid’’ cylinder of radius Rin, at least in the early stage ofreplication (fraction replicated f � 40%). The newly synthesizeddaughter strands, on the other hand, are free to occupy the entirenucleoid volume (of cylindrical radius Rout), which is slightlylarger than its inner-nucleoid compartment. Because they cangain entropy by avoiding contact with the constrained motherstrand, the daughter strands will preferentially favor the periph-eral region of the nucleoid. Below, we show that this simplemodel of the replicating bacterial chromosome captures a num-ber of major features of the experimental data in both E. coli andC. crescentus. We note that, without the concentric-shell model,simulations often show a significant delay in separation ofdaughter strands (bipartitioning along the long axis of the cell)in the early stage of replication (see below), indicating its mainrole in facilitating the onset of segregation.

¶The nucleoid compaction has been well studied and documented in the past half-century.The pioneers in the field are Mason and Powelson (28), who had already observed thecompact nature of bacterial nucleoid in living cell in 1956. For further discussion on othermicroscopic observations and physicochemical considerations, see a recent review byWoldringh and Nanninga (29).

Fig. 3. Chains of nontrivial topology and their typical conformations in strong confinement. (A) A schematic representation of the slowly evolving topologyof circular chromosome during replication. (Note: not to be confused with actual polymer conformations.) (B) Results of two stages of the replication process,indicated by the vertical gray bar in A, namely, the asymmetric and symmetric figure-� conformations, respectively. For each case, a snapshot of a typicalconformation is shown (top of each image) and the spatial distribution of the midpoints of the colored segments (blue, red, and gray) around their averageposition (bottom of each image). (C) The conformation of a figure-eight chain consisting of a total of 2,047 beads (�120 times the width D of the box) in arod-shaped box of aspect ratio 6. Note the mirror-like, linear ordering of the chain within the rod-shaped geometry, although each bead shows a positionalfluctuation comparable with the size of the confinement width D.

12390 � www.pnas.org�cgi�doi�10.1073�pnas.0605305103 Jun and Mulder

Dow

nloa

ded

by g

uest

on

Mar

ch 5

, 202

1

Page 4: Entropy-driven spatial organization of highly confined polymers: … · 2006. 9. 22. · presence of strong confinement. Simulations of two ring poly-mers, in the same setting as

Application to Chromosome Segregation in E. coli. To simulate DNAreplication in slowly growing E. coli, we first created andequilibrated a circular chain inside a cylindrical tube, where atypical initial conformation at the onset of replication is linearwith ori and ter near the opposite cell poles (see the typicalconformation of figure-eight chain above). The chain is ‘‘repli-cated’’ during the course of our simulation by adding a pair ofmonomers at the replication forks each time a fixed number ofMonte Carlo sweeps of the whole system is completed (i.e., ateach fixed ‘‘time’’ interval), mimicking the linear growth in timeof the size of the replication bubble. To take into account thegrowth of the bacterium, we also increased the length of theconcentric shells at the same rate that the replication proceeds.Importantly, we also tested the effect of implementing thepresence of a ‘‘replication factory,’’ which assumes a colocaliza-tion of the bacterial replication machineries (or the two repli-somes, where DNA synthesis takes place) (34, 35). In this case,a figure-eight topology of the replication bubble is imposed byforcing the replication forks to be in each other’s proximity.

In Fig. 5, we present two possible segregation scenarios in theelongating cell of E. coli obtained from our simulations andcompare them with the recent experimental data of Bates andKleckner (7). Together with a series of snapshots of the repli-cating circular chain, we show the trajectories of ori and ter in thepresence (Fig. 5 Left) and in the absence (Fig. 5 Right) of areplication factory. Although one can clearly see that theduplicating chromosomes segregate in both cases, the spatial

distribution of the daughter strands differs significantly in thefirst half of the replication period; when the two replisomes areforced to stay in close proximity to one another, the replicatedDNA adopts a more symmetrical, bimodal distribution alreadymuch before the half-replicated stage than when no such con-straint is imposed. The origin of this early separation of theduplicated ori sites can be easily understood from the repulsionbetween two closed loops, where the replication factory plays therole of a junction (see Spatial Organization of Polymers UnderConfinement), which is facilitated by the outer shell of oursimulated nucleoid. Despite the differences in the degree ofsymmetry between the two cases, we note that the pair of ori‘‘markers’’ almost always appears on opposite sides of the cellcenter, as can be seen by the individual paths (dotted lines) inFig. 5.

On the other hand, ter drifted slowly toward the center of thecell in both cases. As a result of the entropy-driven, globalmovement of both daughter and mother strands, the ori and terpaths cross each other during replication, where on average thecrossing occurs more clearly in the presence of a replicationfactory than in its absence. Interestingly, the experimental resultsby Bates and Kleckner (7) stand between these two extremescenarios (with and without replication factory). Indeed, theDnaX (replisome) markers in their experiment separatedaround the halfway stage of replication, suggesting that repli-somes that are constrained to localize in the cell center duringthe early phases of replication may facilitate the segregation ofdaughter strands (see also Fig. 7, which is published as support-ing information on the PNAS web site, for the case of transientreplication factory confined to the mid-cell).

Application to Chromosome Segregation in C. crescentus and thePrincipal Linear Ordering of Circular Chromosomes in Bacteria. Al-though there are major cell biological differences between C.crescentus and E. coli, the physical aspects of their chromosomessuch as the size and the topology are very similar. Nevertheless,we incorporated one major experimental observation specific toC. crescentus in our simulation, that one of the origins ofreplication (say, ori1) is associated with (36) and stays at the

Fig. 4. The concentric-shell model of replicating nucleoid, which we employin the simulation of DNA replication.

Fig. 5. Chromosome segregation in E. coli, comparing simulation vs. experiment. (Left) A series of typical conformations of a replicating circular chain (motherstrand in gray, two daughter strands in red and blue). We also present two sets of segregation pathways (ori-ter trajectories during replication) in the presence(Left) and absence (Right) of a replication factory. See the schematic diagrams of the chain topology; the black pentagons represent the replisomes. The dottedlines show the results of 10 individual simulations, and the solid lines show average trajectories. (Center) We juxtapose the simulations with the published datain Bates and Kleckner (7) in an attempt to capture the main features of the experimental observations (note that, to show the average cell growth, we have scaledback the normalized cell lengths presented in ref. 7, using their data). For comparison, we used the fraction-replicated f as our ‘‘universal clock’’.

Jun and Mulder PNAS � August 15, 2006 � vol. 103 � no. 33 � 12391

BIO

PHYS

ICS

Dow

nloa

ded

by g

uest

on

Mar

ch 5

, 202

1

Page 5: Entropy-driven spatial organization of highly confined polymers: … · 2006. 9. 22. · presence of strong confinement. Simulations of two ring poly-mers, in the same setting as

stalked pole during replication, as well as that the two replisomesalways remain in the vicinity of each other (3).

In Fig. 6, we show the average (of 26 simulations) trajectoriesof a number of loci on our ‘‘pseudo’’ chromosome and comparethem with the experimental data by Viollier et al. (3). Thesimulation seems to capture the essential features of the spatio-temporal organization of the replicating chromosome in C.crescentus: Newly replicated chromosome loci move rapidly andare deposited sequentially at the cell pole to create the mirrorsymmetry of bacterial chromosome near the end of replication,which is fully consistent with both the principal linear orderingof bacterial chromosome in bacteria [e.g., C. crescentus (3), B.subtilis (37), and E. coli (38, 39)] and with our result on the typicalconformation of figure-eight chain in strong confinement pre-sented above.†† However, we expect considerable cell-to-cellvariations of this linear ordering, because our theoretical pre-dictions indicate that the size of the positional f luctuations ofthe chain segments is comparable with the width of nucleoid (seethe gray area in Fig. 3C). Nevertheless, if specific loci along thechromosome, in particular the ori regions, are tethered to theopposite cell poles after release from the replication factory ashypothesized in the ‘‘extrusion-capture’’ model (41), such fluc-tuations could be reduced (42).

Another issue of interest in C. crescentus is the putative roleof the actin-like MreB proteins in segregation. In their recentwork, Gitai et al. (36) showed that addition of a small moleculeA22, which targets MreB and perturbs its function, blocks themovement of newly replicated loci near the origin of replication.In contrast, addition of A22 after the origins have segregated didnot affect the separation of the origin-distal loci.

In our simulations, we also observed similar failure or signif-icant delay in origin separation when we completely removed theconcentric shells. Thus, because the outer shell facilitates thediffusion and bipartitioning of daughter strands in the early stage

of DNA replication, we favor to interpret the role of MreB inchromosome segregation (if any) as effectively comparable withthat of the outer shell in our concentric-shell model, and viceversa. Indeed, to produce the results in Fig. 6, it was enough toconfine the ori region (�1% of the total chain around ori,corresponding to �400 kb of the C. crescentus chromosome) inthe volume near the stalked pole (36) only until 10–20% of thechain has been replicated, after which point the overall organi-zation established by entropic effects was self-supporting.

Timing and the Concentric-Shell Model. In the two examples aboveof replicating circular chromosomes, we have used the fraction-replicated, f, as our universal ‘‘clock’’ to compare simulation anddata. Although it is difficult to predict theoretically the actualtiming of the segregation process, we can still estimate whetherthe rate of entropy-driven process is consistent with the mea-sured timescale of diffusive motion of DNA segments of bacte-rial chromosome.

Recently, Woldringh’s group (4, 43) tracked a specificallylabeled DNA region using the Lac-O�Lac-I system. The Lacrepressor–GFP fusion protein binds to the DNA section wheretandem repeats of the Lac operator are inserted, which allowedWoldringh and coworkers to monitor the motion of the DNA.The movement of such a GFP spot has been followed both inside(in vivo) and outside (isolated nucleoids) the cell, where themeasured effective diffusion constants are Din � 10�5 �m2�s (4)and Dout � 10�1 �m2�s (43), respectively.‡‡ For the typicallengthscale of �1 �m of a bacterial cell, these diffusion constantsimply a timescale of �out � 10 s and �in � 105 s, respectively, basedon the mean-square displacement relation �x2 � Deff�t� (Deff isthe effective diffusion constant) with � being 1 for ordinarydiffusion. Because the duration of complete duplication of the E.coli chromosome is of order �repl � 103 s, there exists a clearseparation of timescales �out �� �repl �� �in. We interpret thisresult as indicating the following: (i) in the bulk of the relativelystable organization of the in vivo nucleoid, the chromosome doesnot show large-scale diffusion within the timeframe of the cellcycle (�repl � 103 s vs. �in � 105 s), and (ii) in the nucleoidperiphery region, where the DNA is less dense (e.g., in the outershell of our simulated nucleoid), the DNA dynamics could inprinciple be fast enough (�out � 10 s) for the daughter strands tofind their equilibrium conformations during replication (�repl �103 s), at least in the early stage of replication.

Indeed, our concentric-shell–based simulation scenarios, pre-sented in Figs. 5 and 6, reflect this separation of timescales. Inother words, the entropy-driven process can explain the rapidmovement of newly replicated loci (e.g., ori), without requiringany ‘‘mitotic-like’’ apparatus often discussed in the literature (44,11) (see below and also the snapshots in Fig. 5), as well as theslow confined diffusion of ter in the experimental data. Wefurther remark that, in our concentric-shell model, the effectivediffusion rate of the daughter strand can be controlled by thethickness of the outer shell. For example, as the shell becomesthinner than the thickness of the simulated chain, the movementrate decreases, and the ori-ter crossing during replication isgradually delayed (see Fig. 8, which is published as supportinginformation on the PNAS web site). One may think of this effectin terms of increasing ‘‘friction’’ in the outer-shell region.

Future studies should provide more detailed information ofthe dynamics and organization of DNA inside bacteria. Forinstance, it would be highly desirable to see the distribution ofdifferentially labeled daughter strands in the early stage of DNAreplication within the cell.††For E. coli, however, we note that recent data from Sherratt and coworkers (40) show a

tandem-like left–right-arc symmetry of chromosome. Although the significance and theuniversality of their result is yet to be understood by more studies, extension of ourtheoretical approaches should help unravel the cause of similar patterns by furtherconsidering, for example, asymmetric constraints that the leading and lagging strands ofDNA may experience during replication and transcription.

‡‡More recent measurements using fluorescence correlation spectroscopy reveal a fastermotion of DNA segments of isolated nucleoid, where the typical timescale is �1 ms formotion at the scale of �0.2 �m (O. Krichevsky, personal communication).

Fig. 6. Chromosome segregation of C. crescentus, comparing simulation(Left) vs. experiment (3) (Right). The simulated trajectories are the average of26 individual simulation runs (or ‘‘cells’’); we show the trajectories of ninerepresentative loci (including ori and ter) on the right-arc of a circular chro-mosome for the entire duration of replication (up to 99.9%), whereas exper-imental data are only available for trajectories up to 50% of replication. Forclarity, we only show the trajectories from the onset of replication of eachlocus. A full trajectory of ter is shown, however, to emphasize its slow driftfrom the cell pole to the cell center during replication, in contrast to the fast,directed diffusion of ori2 in the nucleoid periphery (on the other hand, wekept ori1 in the volume near the stalked pole until 10–20% of the chain hasbeen replicated). The final spatial organization after the completion of rep-lication bears an interesting resemblance to the mirror-like symmetry of thefigure-eight chain (see Fig. 3).

12392 � www.pnas.org�cgi�doi�10.1073�pnas.0605305103 Jun and Mulder

Dow

nloa

ded

by g

uest

on

Mar

ch 5

, 202

1

Page 6: Entropy-driven spatial organization of highly confined polymers: … · 2006. 9. 22. · presence of strong confinement. Simulations of two ring poly-mers, in the same setting as

Conclusion and OutlookIn this work, we have provided numerical evidence that alltopologically distinct domains of a confined polymer complex ofarbitrary topology effectively repel one another to maximize thetotal conformational entropy. As a result, the conformationalentropy of a duplicating circular chain in a rod-shaped cell byitself provides a purely physical driving force for segregation. Wealso used a simple model of bacterial nucleoid to simulatereplication. Interestingly, this minimal model captures the es-sential features of experimental data on the two model systemsof E. coli and C. crescentus of choice. That a single mechanismcan explain these features in two independent organisms sug-gests that entropic effects should be taken into account whendiscussing the organization and dynamics of the bacterial chro-mosome. It is clear that, in principle, these effects could makemajor contributions to a large number of important chromo-somal phenomena that involve directed movement, either ofindividual loci or of chromosomal domains or even wholechromosomes, from one location to another, quite in line withthe mechanical conception of chromosome organization pro-posed by Kleckner and colleagues (7, 16). These issues can beexplored more fully by considering other prokaryotic species,which occur with a variety of cell shapes and diverse composi-tions of their chromosomes. Finally, it is also tempting tospeculate on the implications of our results on the question of theevolution of DNA segregation mechanisms. It is quite conceiv-able that early life [or the ‘‘First Cell’’ (45)] did not yet possess

sophisticated molecular machineries, so that basic physical prin-ciples such as the maximization of entropy played a prominentrole in its mechanical processes, such as the partitioning ofgenetic materials.

We believe that these are some of the many interestingquestions that could be answered quantitatively by extending theideas and results presented in this work. Moreover, with recentdevelopments in single-molecule manipulation techniques andmicromachined experimental environments such as microchan-nels, our theoretical results are testable in vitro. If indeedvalidated, these considerations might also find applications in thedesign of synthetic cells and minimal organisms.

We thank C. Woldringh for introducing bacterial chromosome segrega-tion to us and for invaluable comments and discussions; D. Bray, R. Th.Dame, O. Krichevsky, A. Lindner, N. Nanninga, S. Tans, and P. R. tenWolde for the critical reading and comments on the manuscript; and D.Frenkel, B.-Y. Ha, and F. Taddei for stimulating discussions. S.J. alsothanks A. Cacciuto and A. Arnold for help with simulation and figures,J. Raoul-Duval for inspiring conversations on entropy, and P. McGrath(Shapiro and McAdams Laboratories, Stanford University, Stanford,CA) and D. Bates (Kleckner Laboratory, Harvard University, Cam-bridge, MA) for sending the data published in refs. 3 and 7. S.J. wassupported by a Natural Sciences and Engineering Research Council(Canada) postdoctoral fellowship. This work is part of the researchprogram of the Stichting voor Fundamenteel Onderzoek der Materie(FOM), which is supported by the Nederlandse organisatie voor Weten-schappelijk Onderzoek (NWO).

1. Jacob, F., Brenner, S. & Cuzin, F. (1963) Cold Spring Harbor Symp. Quant. Biol.23, 329–348.

2. Webb, C. D., Graumann, P. L., Kahana, J. A., Teleman, A .A., Silver, P. A. &Losick, R. (1998) Mol. Microbiol. 28, 883–892.

3. Viollier, P. H., Thanbichler, M., McGrath, P. T., West, L., Meewan, M.,McAdams, H. H. & Shapiro, L. (2004) Proc. Natl. Acad. Sci. USA 101,9257–9262.

4. Elmore, S., Muller, M., Vischer, N., Odijk, T. & Woldringh, C. L. (2005) J.Struct. Biol. 151, 272–287.

5. Webb, C. D., Teleman, A., Gordon, S., Straight, A., Belmont, A., Lin, D. C. H.,Grossman, A. D., Wright, A. & Losick, R. (1997) Cell 88, 667–674.

6. Niki, H. & Hiraga, S. (1998) Genes Dev. 12, 1036–1045.7. Bates, D. & Kleckner, N. (2005) Cell 121, 899–911.8. Woldringh, C. L. (2002) Mol. Microbiol. 45, 17–29.9. Errington, J., Murray, H. & Wu, L. J. (2005) Philos. Trans. R. Soc. London B

360, 497–505.10. Gitai, Z., Thanbichler, M. & Shapiro, L. (2005) Trends Microbiol. 13, 221–228.11. Gitai, Z. (2005) Cell 120, 577–586.12. Gerdes, K., Moller-Jensen, J., Ebersbach, G., Kruse, T. & Nordstrom, K. (2004)

Cell 116, 359–366.13. Sherratt, D. J. (2003) Science 301, 780–785.14. Marko, J. F. & Siggia, E. D. (1997) Mol. Biol. Cell 8, 2217–2231.15. Trun, N. J. & Marko, J. F. (1998) ASM News 64, 276–283.16. Kleckner, N., Zickler, D., Jones, G. F., Dekker, J., Padmore, R., Henle, J. &

Hutchinson, J. (2004) Proc. Natl. Acad. Sci. USA 101, 12592–12597.17. de Gennes, P. G. (1997) Scaling Concepts in Polymer Physics (Cornell Univ.

Press, Ithaca, NY).18. Flory, P. J. & Krigbaum, W. R. (1950) J. Chem. Phys. 18, 1086–1094.19. des Cloizeaux, J. (1975) J. Phys. 36, 281–291.20. Daoud, M., Cotton, J. P., Farnoux, B., Jannink, G., Sarma, G., Benoit, H.,

Duplessix, C., Picot, C. & de Gennes, P. G. (1975) Macromolecules 8, 804–818.21. Grosberg, A. Y., Khalatur, P. G. & Khokhlov, A. R. (1982) Makromol. Chem.

Rapid Commun. 3, 709–713.22. Kruger, B., Schafer, L. & Baumgartner, A. (1989) J. Phys. 50, 3191–3223.

23. Dautenhahn, J. & Hall, C. K. (1994) Macromolecules 27, 5399–5412.24. Louis, A. A., Bolhuis, P. G., Hansen, J. P. & Meijer, E. J. (2000) Phys. Rev. Lett.

85, 2522–2525.25. Likos, C. N. (2001) Phys. Rep. 348, 267–439.26. Woldringh, C. L. & Odijk, T. (1999) in Organization of the Prokaryotic Genome,

ed. Charlebois, R. L. (Am. Soc. for Microbiol. Press, Washington, DC), pp.161–197.

27. Cates, M. E. & Deutsch, J. M. (1986) J. Phys. 47, 2121–2128.28. Mason, D. J. & Powelson, D. M. (1956) J. Bacteriol. 71, 474–479.29. Woldringh, C. L. & Nanninga, N. (2006) J. Struct. Biol., 10.1016�

j.jsb.2006.04.013.30. Dame, R. T. (2005) Mol. Microbiol. 56, 858–870.31. Odijk, T. (1998) Biophys Chem. 73, 23–30.32. Sawitzke, J. A. & Austin, S. (2000) Proc. Natl. Acad. Sci. USA 97, 1671–1676.33. Holmes, V. F. & Cozzarelli, N. R. (2000) Proc. Natl. Acad. Sci. USA 97,

1322–1324.34. Lemon, K. P. & Grossman, A. D. (1998) Science 282, 1516–1519.35. Dworkin, J. & Losick, R. (2002) Proc. Natl. Acad. Sci. USA 99, 14089–14094.36. Gitai, Z., Dye, N. A., Reisennauer, A., Wachi, M. & Shapiro, L. (2005) Cell 120,

329–341.37. Teleman, A. A., Graumann, P. L., Lin, D. C. H., Grossman, A. D. & Losick,

R. (1998) Curr. Biol. 8, 1102–1109.38. Niki, H., Yamaichi, Y. & Hiraga, S. (2000) Genes Dev. 14, 212–223.39. Roos, M., van Geel, A. B. M., Aarsman, M. E. G., Veuskens J. T. M.,

Woldringh, C. L. & Nanninga, N. (2001) Mol. Microbiol. 39, 633–640.40. Wang, X., Liu, X, Possoz, C. & Sherratt, D. J. (2006) Genes Dev. 20, 1727–1731.41. Lemon, K. P. & Grossman, A. D. (2001) Genes Dev. 15, 2031–2041.42. Breier, A. M. & Cozzarelli, N. R. (2004) Proc. Natl. Acad. Sci. USA 101,

9175–9176.43. Cunha, S., Woldringh, C. L. & Odijk, T. (2005) J. Struct. Biol. 150, 226–232.44. Moller-Jensen, J., Jensen, R. B., Lowe, J. & Gerdes, K. (2002) EMBO J. 21,

3119–3127.45. Koch, A. L. (2001) Bacterial Growth and Form (Kluwer Academic, Dordrecht,

The Netherlands).

Jun and Mulder PNAS � August 15, 2006 � vol. 103 � no. 33 � 12393

BIO

PHYS

ICS

Dow

nloa

ded

by g

uest

on

Mar

ch 5

, 202

1