Christoph
Schneck
ab,
Jan
Smrek†
a,
Christos N.
Likos
*a and
Andreas
Zöttl†
a
aFaculty of Physics, University of Vienna, Boltzmanngasse 5, 1090 Vienna, Austria. E-mail: christoph.schneck@ehu.eus; jan.smrek@univie.ac.at; christos.likos@univie.ac.at; andreas.zoettl@univie.ac.at
bCentro de Física de Materiales (CSIC, UPV/EHU) and Materials Physics Center MPC, Paseo Manuel de Lardizabal 5, 20018 San Sebastián, Spain
First published on 28th March 2024
We apply monomer-resolved computer simulations of supercoiled ring polymers under shear, taking full account of the hydrodynamic interactions, accompanied, in parallel, by simulations in which these are switched off. The combination of bending and torsional rigidities inherent in these polymers, in conjunction with hydrodynamics, has a profound impact on their flow properties. In contrast to their flexible counterparts, which dramatically deform and inflate under shear [Liebetreu et al., Commun. Mater. 2020, 1, 4], supercoiled rings undergo only weak changes in their overall shape and they display both a reduced propensity to tumbling (at fixed Weissenberg number) and a much stronger orientational resistance with respect to their flexible counterparts. In the presence of hydrodynamic interactions, the coupling of the polymer to solvent flow is capable of bringing about a topological transformation of writhe to twist at strong shear upon conservation of the overall linking number.
Besides the materials perspective, the properties of non-equilibrium supercoiled DNA polymers touch upon two other highly active research areas. First, the circular DNA is found naturally in the form of (typically supercoiled) bacterial plasmids,10 extrachromosomal DNA of eukaryotes,11 kinetoplast DNA12,13 of trypanosoma and supercoiled segments have been suspected to form distinct architectural and functional domains also on chromosomal DNA.14 Supercoiling itself can affect gene expression15 or DNA metabolism16 by modulating access to distinct regions. In the biological context the DNA molecules are also typically out of equilibrium, subject to flows and stresses arising through the action of molecular motors and inducing non-equilibrium conformations17 and dynamics18 that in turn can impact the biological function.19
Second, the properties of unknotted ring polymers under various conditions are one of the largest unsolved problems in polymer physics. The crux lies in the difficulty of capturing the restrictions that the fixed ring topology imposes on the phase space in the presence of other interactions. This general principle is present across different concentration regimes. In dense many-chain systems the chain topology prevents mutual ring linking, resulting in compact, fractal tree-like, conformations,20,21 power-law stress relaxation,22 distinct shear-thinning exponents,23 extensional viscosity thickening,24,25 viscosity of ring-linear blend exceeding that of the two components26,27 or vitrification upon heterogeneous activation.28 Some of these effects are due to ring interpenetration (threading) by other chains, although their general impact on the viscoelastic properties is still an open question.29–32 However, the impact of threadings is apparent in simulations and experiments on supercoiled plasmids. There, the supercoiling induces tighter branched ring conformations, reduces mutual ring interpenetration and speeds up the dynamics.5 Therefore, investigating the systems with torsional degrees of freedom can help to understand dynamical effects also in torsionally relaxed rings, by giving a control on the degrees of the branching and the threading. In semi-dilute and dilute conditions, both, the interchain interactions and the hydrodynamics coupled with the ring topology of the chains affect the chain relaxation dynamics and fluctuations of the metric properties.33–35 For dilute solutions under non-equilibrium flow conditions, the combination of topology with hydrodynamic interactions has dramatic effects on the conformations and dynamics of the solutes. Vorticity swelling in shear,36 extensional or mixed flows37,38 has been predicted in simulations and recently confirmed experimentally,39 whereas hydrodynamic inflation under steady shear40 results into a strong suppression of tumbling and of the Brownian nature of the polymer dynamics for certain domains of shear rate for flexible polymers. To summarize, the supercoiling reduces penetrable area of rings which is relevant for viscoelasticity, but hydrodynamic swelling acts in the opposite way. Therefore here we aim to study systematically in simulation how does the hydrodynamic interaction couple with the supercoiled ring structure in dilute (single chain) conditions in shear. To do so, we employ the multiparticle collision dynamics to account for the hydrodynamic interaction in combination with molecular dynamics for the time evolution of the polymer.
Under non-equilibrium conditions such as shear flow that we focus on in here,40,41 non-local hydrodynamic effects play an essential role for the transport and dynamical properties of polymers. Considering the molecular dynamics of all of the fluid particles interacting with the polymers would be too time-consuming on the typical experimental time-scales (seconds). Hence, several coarse-grained simulation techniques had been developed which capture hydrodynamic interactions in polymers, such as dissipative particle dynamics (DPD),42 the lattice Boltzmann method (LBM)43 and multiparticle collision dynamics (MPCD).44 In particular MPCD had been heavily used in the last two decades to study polymer dynamics under equilibrium and under shear flow conditions. In MPCD the Newtonian background solvent is modeled by effective point-like fluid particles which perform alternating streaming and collision steps.44,45 While intramolecular forces are advanced through molecular dynamics (MD) purely between monomers, the hydrodynamic effects on the monomers stem from monomer-fluid particle interactions. In the simplest and most popular approach of such a hybrid MPCD-MD scheme the monomers are assumed to be pointlike but heavier than the fluid particles and exchange momentum with the local fluid particles in the collision step.46 This allows simulations of single/dilute47–49 and semidilute/dense50,51 polymer solutions in the presence of an explicit solvent. Simple unidirectional shear flow is usually realized by applying Lees–Edwards boundary conditions,47,48,52,53 and the flow can be disturbed by the presence of polymers and their coupling to the fluid motion.
MPCD had been used to quantify the shear-thinning behavior of linear flexible,48,51 semiflexible54 and stiff47,50 polymer solutions, as well as the dynamics of single crosslinked polymer chains.55,56 Furthermore, the MPCD method can easily be adapted to simulate pressure-driven Poiseuille flow bounded by no-slip channel walls.57–59 Consequently the dynamics of single linear flexible60 and semiflexible61 polymers, as well as the shape transitions of a tethered linear polymer62 had been investigated. The properties of polymers with other architectures under shear flow had also been studied in detail, such as star polymers and dendrimers,41,49,63–68 as well as ring polymers.36,40,69,70 Noteworthy, in MPCD hydrodynamic interactions can be turned off easily without modifying other physical properties of the fluid.40,41 This allows to directly determine the influence of hydrodynamic interactions in polymer simulations, of particular importance in out-of-equilibrium conditions.71–73
The article is organized as follows: in section 2 we introduce the model and in section 3 the quantities that characterize the topology and the shapes of the supercoiled polymers in equilibrium and under shear. In sections 4 and 5 we present the extracted equilibrium properties and the behavior under shear flow of supercoiled ring polymers, respectively. We conclude our work in section 6, whereas certain technical details of our work are delegated to the Appendix.
(1) |
Periodic boundary conditions are applied for particles crossing the dimensions of the simulation box. In the subsequent collision step the entire simulation box is coarse-grained into cubic cells of a grid with lattice constant ac and the solvent particles are sorted according to their current positions. The velocities of all Nc particles in one particular cell are summed up to a center-of-mass velocity, , and a rotation matrix R[α] is generated and assigned to the cell. The orientation of the rotation axis is picked randomly at each step and varies from cell to cell, whereas the rotation angle α is fixed. This randomness is the ingredient endowing MPCD with thermal fluctuations. The velocities of all particles within the cell are then rotated relative to the center-of-mass velocity of the cell,
(2) |
There are three basic units all set to unity: the mass ms of the fluid particles, the length of the cubic collision cells ac, and the energy kBT specifying the average kinetic energy of a solvent particle by . This corresponds to expressing length, energy, and mass in units of ac, kBT, and ms. A reference unit for time can then be derived as . The box dimensions are set to Lx = 80ac, Ly = 50ac, Lz = 50ac, which has been shown to be large enough to avoid flow interference through periodic boundary conditions or polymer self-interactions for similar simulations.40 The box is rectangular with its long side in the shear flow direction – see later below the description of the system. The time between collisions is set to h = 0.1τMPCD, the collision angle α = 130° and the solvent particle number density , where 〈Nc〉 = 10 is the average number of solvent particles in each collision cell. This choice of parameters corresponds to a dynamic viscosity75 of η = 8.7τMPCD−1msac−1.
For short time steps h the mean free path between consecutive collisions is small compared to the cell size. The discrete nature of the grid causes particles to stay in the same cell for several collisions before particle exchange with neighbouring cells occurs. This causes particles within one collision cell to build up correlations and breaks the translational symmetry of the system. Galilei invariance can be recovered by performing a random shift of the sorting grid by the shift vector with the components that are sampled uniformly from . Additionally, the shifting procedure accelerates momentum transfer between particles.76
To generate shear flow with a tunable shear rate the periodic boundaries are modified to Lees–Edwards boundary conditions.52 This defines three distinct space directions: the shear motion is along the flow direction (x-axis), the magnitude of shear flow varies along the gradient direction (y-axis), and the remaining one is called the vorticity direction (z-axis). Thereby, energy is constantly pumped into the system, and then is converted into heat due to the MPCD-fluid viscosity, which, if left unattended, would result into constant viscous heating of the system. A thermostat is thus necessary in such a non-equilibrium system to maintain the temperature at its prescribed value. We employ here a cell-level thermostat,44,77 which further ensures that the particle velocities follow a Maxwell–Boltzmann distribution and it restricts the fluctuations of the total energy of the system to those leading to the desired value kBT.
To create coupling with the fluid and thereby enable HI, we include the monomers in the MPCD collision, step46 but other approaches also exist.78 An embedded monomer i with mass mm and velocity contributes to the center-of-mass velocity of the collision cell. If Nm is the number of monomers in a particular cell, the center-of-mass velocity is
(3) |
A stochastic rotation as in eqn (2) is then performed on the relative velocities of both the solvent particles and the embedded monomers. This results in an exchange of momentum and energy between all particles within a collision cell. The updated monomer velocities are then used as initial conditions for the subsequent MD time step. 100 MD time steps are performed between consecutive MPCD collisions.
Gompper et al.44 have noted that the average number of monomers per collision cell should be smaller than unity in order to properly resolve HI between them. The bond length of polymer chains should be of order of the cell size ac so that monomers are close enough to display effects due to HI but far enough apart to avoid multiple monomers within a cell. Furthermore, Ripoll et al.79 have demonstrated that the mass ratio mm/ms should equal the average number of solvent particles per cell 〈Nc〉 (assuming there is only one monomer per cell). Under this condition the monomer collides with a piece of solvent with equal mass so that the mutual momentum exchange is balanced. For mm ≫ 〈Nc〉ms the polymer is barely affected by the solvent, whereas for mm ≪ 〈Nc〉ms the monomers are kicked around too violently. Hence we chose mm = 〈Nc〉ms = 10ms.
In order to study the specific effects of the hydrodynamic interactions on the dynamics of the system the same simulations have to be performed with and without HI. An alternative concept of simulating the solvent and the interaction with the monomer has to be found, that differs as little as possible from the original conditions except for the presence of HI. Ripoll et al. invented an efficient method to switch off HI in a MPCD algorithm.80 They suggest that the positions of the solvent particles do not contribute in the collision step and therefore, no explicit particles have to be considered at all. Instead, every monomer is coupled with an effective solvent momentum that is chosen randomly from a Maxwell–Boltzmann distribution of variance ms〈Nc〉kBT and zero mean. A center-of-mass velocity is assigned to every monomer velocity given by:
(4) |
(5) |
Neighbouring beads are attached to each other along the backbone by applying a pair-wise attractive FENE (finitely extensible nonlinear elastic) potential given by
(6) |
We further include bending and torsional stiffness as in Smrek et al.,5 to mimic the coarse-grained properties of the DNA. We show a sketch of the polymer in Fig. 1(a), and we describe the model in more detail in what follows. The bending energy is introduced between every consecutive triplets of chain neighbours , , and (again respecting the ring topology):
Ubend(θb) = kbend(1 − cosθb), | (7) |
(8) |
The bending constant is chosen to be kbend = 20kBT leading to a persistence length lp = 20σ. The scale of the monomer diameter should be mapped5 to DNA's thickness in real units of σ = 2.5 nm, i.e., 2.5/0.34 = 7.35 bp per bead and lp = 150 bp = 50 nm.
To model the torsional degrees of freedom, every bead is equipped with three patches which remain at a fixed distance . The vectors are defined as the vectors going from the position of bead i to its k-th patch where k = 1, 2, 3 so that each triplet forms an orthogonal basis with the respective bead position as the origin. In Fig. 1(a), we show a short segment of a polymer chain with patches where colours blue, red, and green correspond to the numeration k = 1, 2, 3, respectively. The blue and red patches are used to introduce two dihedral springs that constrain the torsional angle ψb/r between consecutive beads close to a fixed equilibrium value ψ0. Here, ψb/r is defined as the angle between the planes with normal vectors and , see Fig. 1(c), and it is evaluated for the blue and red patches, respectively. The corresponding potential is given by the expression:
(9) |
Ualign(θa) = kalign(1 − cosθa), | (10) |
(11) |
In this way, blue and red patches are aligned orthogonal to the polymer backbone consisting of the beads so that dihedral angles between consecutive beads can be formed properly, see Fig. 1(d). The constant is chosen to be as high as kalign = 200kBT to minimize the fluctuations of the alignment.
Forces derived from all potentials presented act on the beads i whereas torsion, eqn (9), and alignment, eqn (10), also yield forces upon the respective patches k denoted by . Forces acting on the beads are calculated as
(12) |
(13) |
(14) |
Adding patches to the beads effectively transforms the previous point particles i to rigid bodies with a rotational degree of freedom χi, angular velocity , and inertia tensor Iiαβ. The set of three angles χi defines the orientation of the monomer i with respect to the lab frame, and hence define the positions of the patches on the monomer i. Their explicit representation with quaternions is discussed in the Appendix. The monomers are modeled as extended spheres with radius lpat and continuous mass density ρm = 3mm/(4πlpat3), which corresponds to a trivial inertia tensor Iiαβ = Isphereδα,β with the moment of inertia . The position of the bead is the central point of the rigid body and coincides with the center-of-mass of the monomer, whereas the patches reside on the sphere's surface. Therefore, only the forces contribute to the total torque on monomer i, expressed as
(15) |
(16) |
where and are points on the curve Ct and at a given time t. The writhe gives an intuitive measure of how many crossings an average 2D projection of the molecular backbone forms with itself. For the actual computation we use a version of eqn (16) detailed in Klenin and Langowski,81 where the bead positions serve as the reference points of the discretized curve Ct.
The twist is computed for the torsionally constrained polymer rings by summing over the difference of all dihedral angles from the respective equilibrium value ψ0:
(17) |
The blue and red patches each form an individual set of dihedral angles , respectively which is why an arithmetic mean over the two is taken in the evaluation of in eqn (17). The alignment potential forces the green patches to be orientated nearly perfectly with the polymer backbone so that the blue and red patches are kept in a perpendicular orientation to it. A high value is picked for the constant kalign = 200kBT which limits the fluctuations from this structure. For this reason, the vectors serve as (unnormalized) normal vectors to the space curve approximated by the bead positions . With the brackets 〈…〉t denoting averaging over time, we also define the quantities:
(18) |
The patches endow the simulated ring polymer with the structure of a closed ribbon, which therefore obeys Călugăreanu's theorem, stating that the sum over its writhe and twist, i.e., its linking number Lk is a conserved quantity:82
(19) |
The linking number is thus a topological invariant of the ribbon, independently of time or applied shear rate, akin to the knotedness of single rings or the catenation connectivities of several ones in composite macromolecular entities, such as catenanes. Note that the model is torsionally symmetric, i.e., it does not distinguish between left-handed and right-handed torsion and therefore we have chosen one of them and we do not specify the sign of the supercoiling or the linking number.
Supercoiled rings must be initialized in a state with minimal torsional energy to avoid large initial torsional forces that could cause a simulation crash. The initial conformations of rings with σsc > 0 are generated by starting from a flat open ribbon with a linking number of zero and performing MD steps while gradually ramping up the offset of the torsional potential ψ0 from zero to the desired angle 2πσsc in small steps. During this procedure monomers are coupled to the random MPCD solvent. The offset is increased in steps of 10–3 × 2πσsc after every 2000th collision with the solvent until the final angle is reached. The resulting conformations are supercoiled rings for which is indeed conserved up to fluctuations around the mean. The initialization process is performed for the supercoiling values σsc = 0.01 and σsc = 0.02. In Fig. 2, we show snapshots of all four types of ring polymers investigated.
Fig. 2 Snapshots of the supercoiled polymers at equilibrium . (a) Relaxed ring with ktorsion = 0, (b) σsc = 0.00, (c) σsc = 0.01, (d) σsc = 0.02. |
(20) |
and additional quantities derived from it, to be specified in what follows. Here, is the α-coordinate of i-th monomer's position relative to the center-of-mass of the polymer at a given time t and shear rate . The diagonal element reflects the elongation of the polymer along the α-coordinate. Of particular importance are the instantaneous eigenvalues , out of which several useful rotational invariants can be constructed. The instantaneous radius of gyration squared, , is defined as the trace of the gyration tensor:
(21) |
(22) |
The shape of the conformation can be characterized by the prolateness , defined as:40
(23) |
(24) |
where . For this quantity, occurs if all eigenvalues are identical (e.g. spherical or polyhedral group symmetry) while indicates that two eigenvalues are zero (e.g. all monomers on a line). The (squared) gyration radius, the prolateness and the relative shape anisotropy are all independent of the orientation of the polymer in space. Similar to the writhe and twist, eqn (18), we define time-averages as
(25) |
where = Gαβ, Rg2, Δx, S*, or δ*, and we report on their values in what follows.
For the supercoiled rings with nonzero Linking number, σsc > 0, the simulation occasionally crashed, due to the relatively large timestep and the buildup of numerical instabilities leading to unphysically large forces. In such cases, we discarded the data. Higher degrees of writhe constrain the rings to take on more contracted conformations and the resulting contorted macrostructure of the polymer forces some beads to be pressed together increasing the risk of building up large repulsive forces. High shear rates amplify this behaviour as the solvent flow stretches out the polymer rings. More specifically, two different kinds of undesired behaviour have been observed. The first is rupture of the polymer, where a small part (usually up to 5 monomers) leaves its place within the chain. The second type of undesired behaviour amounts to Lk spontaneously being changed by an integer value ±1, violating the topological constraints that force the ring to have a constant linking number (apart from small deviations due to the fact that the patches are not necessarily aligned in a perfect right angle relative to the backbone). These crashes happened mostly when using the −HI-solvent at high shear rates, since the total number of MD steps was much higher compared to the costly particle-based solvent with +HI and the chances of a crashing situation increased with the length of the simulation runs which could not be decreased significantly. Both types of crashes can presumably be avoided by using a smaller ΔtMD.
In Fig. 3(a)–(d), we show time series of the topological quantities, writhe and twist, for all systems simulated, whereby the curves of the single runs are averaged to a single curve. Whereas both the relaxed rings and the rings without supercoiling assume conformations that are essentially writhe-free, the rings with σsc = 0.01 and σsc = 0.02 display conformations that carry finite writhe. Connected to this, and rather insensitively of the degree of supercoiling, the twist fluctuates closely around zero in case of σsc = 0.00 whereas for σsc > 0, takes on small but nonzero values. Up to this value, the writhe equals the linking number for the entire duration of the simulation and Călugăreanu's theorem (19) is indeed fulfilled for all systems. The relaxed chains remain in a state with only low writhe because of the high bending stiffness-cost that a high writhe would incur, and the twist fluctuates arbitrarily for the same chains because of the absence of torsional springs. The overall behaviour of the topological parameters does not depend on the presence of HI, as it should be the case for polymers in equilibrium.
The equilibrium distributions of the squared radius of gyration are presented in Fig. 3(e) and the expectation values of this quantity are summarized in Table 1, which also provides a comparison with the gyration radius of a fully flexible ring. As a first remark, we note that the presence of the bending stiffness, independently of the supercoiling, causes the rings to swell with respect to their flexible counterparts, as expected. For σsc = 0.00 both the bending and the torsional stiffness force the ring to remain in an open circular shape which leads to a sharp peak in the Rg2(t;0)-distributions. The absence of torsional stiffness softens considerably the relaxed rings, leading to slightly smaller sizes as well as a broader distribution of the gyration radius values. The σsc = 0.00 rings are locked in a rather open and flat conformation and Rg2(t;0) fluctuates more weakly around a larger mean compared to the relaxed counterparts. The values of the gyration radius agree with Smrek et al.,5 which confirming the correct implementation of the model. Moreover the Rg-distribution of the relaxed case shows a long tail at low values, indicating a weakly bimodal character of the distribution in agreement with the findings of Smrek et al.5 Nonzero supercoiling forces the polymers to take on more complex supercoiled shapes, characterized by a smaller Rg2(t;0) for σsc > 0, as can be also confirmed by the prolateness data below. The values of Rg2(t;0) coincide for the solvents with and without HI, again as expected, and confirming that equilibrium has indeed been achieved for both simulation variants.
The equilibrium distributions of the prolateness parameter S* are depicted in Fig. 3(f). As intuitively expected, the σsc = 0.00-ring and the relaxed one are mainly found in oblate shapes as they tend to be open rings, a property manifested in the fact that their S*(t;0)-distributions peak at values below zero. Nevertheless, even for these rings the probability distributions have sufficiently long tails into the prolate domain, resulting into expectation values S*(0) ≅ 0 overall, see Fig. 7(d) (at ): to obtain rings that are oblate on average, one would need a higher value of kbend and/or a smaller value of N, so as to obtain a larger ratio of persistence to contour lengths. On the other hand, supercoiling increases the prolateness markedly, as it constrains the rings in a more elongated conformation, consistent with the increasingly writhed conformations they assume. Accordingly, the prolateness distribution of the σsc = 0.01-supercoiled ring has a peak at S*(t, 0) ≅ 0.5 and that of the σsc = 0.02-ring at S*(t,0) ≅ 1.25. Despite the increase in prolateness, the gyration radius of the σsc = 0.02-ring remains unchanged with respect to that of its σsc = 0.01-counterpart, as the elongation in one direction is compensated by shrinking in the two orthogonal ones.
We begin our analysis by looking at the topological characteristics of the supercoiled rings, and in particular at the writhe and the twist , reported in Fig. 4(a)–(d). The linking number is also depicted as their sum, which is constant for all , serving as a test of the validity of the simulations. In all cases and for most shear rates considered, writhe and twist remain constant at their preferred values as in equilibrium according to their degree of supercoiling σsc; hydrodynamic forces are not capable of “unwrithing” the molecules for most of the range of shear rates considered. However, for the highest values of tested and in the +HI-variant, where simulations at these shear rates could be performed without crashes, considerable transformation of into takes place (see Fig. 4(a) and (c) at ). This is interpreted as a forced opening of the writhed structure since the process is connected with a high torsional energy penalty. The snapshots of supercoiled rings with σsc = 0.02 and σsc = 0.01 in Fig. 4(e) and (f), respectively (both +HI), are representatives of the polymers’ behaviour under shear flow in comparison to their shapes in equilibrium . At zero shear rate the σsc = 0.02-ring rotates freely in space and takes on more compact, prolate shapes. Although bending and torsion constrains it to have a writhe of nearly it is not elongated in any direction. In contrast, the polymer is almost completely stretched out and aligned along flow direction at high shear rate of . Above the shear flow is even able to counteract the torsional forces so that writhe and twist experience large fluctuations and the conformation unwrithes, see Fig. 4(e) and (f).
In Fig. 5 we show the diagonal elements of for both cases when hydrodynamic interactions are active (+HI) as well as the situation when the polymer is coupled to the random MPCD solvent (−HI). All ring polymers considered in this work are not affected by the shear flow up to a certain threshold shear rate of for + HI and for −HI. As of the behaviour of polymers under shear flow starts deviating from its equilibrium pattern at Weissenberg number Wi× ≅ 1, this suggests that the longest relaxation times should be of the order τR ≅ 104 for +HI and τR ≅ 105 for −HI, in agreement with previous findings for flexible ring polymers36 or star polymers,66 where the two relaxation times were indeed found to differ by about one order of magnitude for polymers of N ≅ 100.‡
For all degrees of supercoiling as well as for the relaxed rings the component of the gyration tensor in the flow direction grows with increasing shear rates while the components in vorticity and gradient directions both get smaller. As the curves do not differ between relaxed and supercoiled rings, we can attribute the trends as well as their deviations from those of flexible rings to the non-vanishing bending rigidity of the polymers. Interestingly, the growth of (i.e. size extension in the flow direction) with shear rate, shown in Fig. 5(a), is moderate limited to a factor of ≈2–3 with respect to its equilibrium value for all shear rates covered, which is small compared to the growth factor of the fully flexible rings,36,40 which reach values up to ≈15–20. The rings at hand do not stretch as much in the flow direction as the flexible ones, due to the stiffness of the bonds that prevents elongated conformations, which would necessarily result into sharp turns and thus very high bending penalties at the tips of the rings. For the −HI case, Fig. 5(b) there is even a saturation of at the highest shear rates, corresponding to 102 ≲ Wi ≲ 103. In Fig. 6, we show a quantity closely related to , namely the extent . The growth of the extent over its equilibrium value is by about a factor 2 for the whole range of shear rates applied, much less than values experimentally measured for fully flexible rings under similar conditions in, e.g., extensional flow.84
Fig. 6 The extent of the polymer along the flow direction normalized over its equilibrium value, Δx(0), as a function of the shear rate . (a) +HI-case; (b) −HI-case. |
In the gradient and vorticity directions (y and z directions respectively), the diagonal elements of the gyration tensor decrease with shear rate but again in ways that differ from those of the flexible cyclic polymers. The contraction in gradient direction is more pronounced than in vorticity direction, as is also the case for flexible rings, but here it decreases with a stronger power-law, as opposed to the power-law , valid for flexible rings,36 independently of the inclusion or exclusion of HI, see Fig. 5(c) and (d). One would naively expect that this is due to a stronger alignment of the rigid rings with the flow direction but, as we will establish in what follows, this is not the case. The rigid and supercoiled rings at hand align, in fact, much less with the flow direction than flexible ones but at the same time they also tumble much less. Accordingly, time intervals of tumbling, in which the (flexible) rings extend stronger into the gradient direction contribute much less to the time-averages and the contraction of the gradient-direction extent with the shear rate is less strong for the flexible rings, resulting into new exponents of the semiflexible ones.
The vorticity-direction dependence of the gyration tensor, shown in Fig. 5(e) and (f), is also quite unique for the rigid rings. The phenomenon of vorticity swelling of the flexible rings36,40 is absent here, because the solvent back-flow is not strong enough to stretch the stiff bonds of these rings. One sees, nevertheless, the solvent backflow effect in comparing the +HI-case, Fig. 5(e), in which it is present, with the −HI-case, Fig. 5(f), in which it is absent. Indeed, in the latter the ring contracts much more in the vorticity direction than in the former, where solvent backflow creates additional swelling pressure in the ring's interior.36 In this respect, it is important to notice that the conversion of writhe to twist and the reduction in vorticity contraction are both effects that take place solely in the presence of HI. Accordingly, we claim that the solvent backflow is crucial for both, i.e., only solvent backflow generates the necessary local stresses along the polymer backbone to bring about the topological conversion while at the same time keeping the ring sufficiently open in the vorticity direction, consistently with the unwrithed conformation seen, e.g., in Fig. 4(f).
The dependence of the shape parameters on the shear rate is summarized in Fig. 7. The decrease of and is compensated by the increase in the flow direction, so that is more or less constant over most of the range of , see Fig. 7(a) and (b). This is again in stark contrast with flexible rings, for which is dominated by and it grows rapidly with shear rate.36,40 Here, in fact, the radius of gyration of all ring types decreases by a small amount before experiencing growth again for very high shear rates. The effect is observed more clearly in the case of −HI because the sample size covers longer time periods.
For the overall polymer size, as expressed by the gyration radius, neither the degree of supercoiling nor the hydrodynamics seem to significantly affect its dependence on shear rate. This changes if one looks at more detailed shape characteristics, such as the prolateness, Fig. 7(c) and (d), and anisotropy, Fig. 7(e) and (f). As they share common characteristics, we discuss and together. First of all, for the lower shear rates, there is now a clear splitting of the curves depending on the degree of supercoiling. Whereas the relaxed ring as well as the σsc = 0.00 start with and low values at low shear rates, which grow as increases, the supercoiled rings are already prolate and very anisotropic at equilibrium, due to the writhed conformations they assume, see Fig. 3. Accordingly, the prolateness and shape anisotropy are affected very weakly by shear for the rings with high supercoiling degree, especially for the +HI-case, whereas in the −HI case prolateness and anisotropy become monotonically increasing at high and all rings follow the same curves there. As mentioned above, solvent backflow for the +HI-case reduces prolateness and anisotropy by temporarily stabilizing more open, oblate conformations of the rings and by converting writhe into twist for the supercoiled ones.
The average alignment of a polymer with the shear flow direction and its dependence on shear rate can be quantified by the orientational resistance , a dimensionless quantity that can be constructed from the gyration tensor elements as:85
(26) |
The strong orientational resistance of the supercoiled polymers, in conjunction with the previous results on the dependence of on the shear rate, shown in Fig. 5, creates an apparent contradiction. Indeed, if we visualize the polymer as an object oriented on a stable configuration that forms an angle with the x-axis, and making the small-angle approximation , valid only for the −HI-case, we can write:
(27) |
Ignoring the weak dependence of on the shear rate [Fig. 5(b)] (fitted exponent 0.03, for , not shown) and writing , eqn (27) yields the power-law , stronger than the measured power-law in Fig. 5(c). Moreover, the corresponding power-laws for flexible polymers read and , i.e., in that case the gradient-direction contraction is weaker than that of the rigid rings although their alignment with the flow direction is stronger. This can be partially explained by the fact that for the flexible ring case, ; even so, it is evident that the picture of an effective ellipsoid orienting itself at a stable angle with the x-axis during shear flow is too simplistic, as the polymer dynamics is richer in patterns developed by the combination of thermal and hydrodynamic forces on the monomers.
Polymers have been shown to undergo tumbling motion under shear flow both in experiment86,87 and simulations.36,69 The tumbling dynamics are characterized by a rapid collapse in one direction (x-axis) followed by the expansion in another direction (y-axis). Such events are cyclical yet not periodic, as the molecule undergoes intermittent phases between stable orientations and vivid tumbling motions. During the latter, a tumbling frequency can be deducted from the cross-correlation function between fluctuations of and around their mean values separated by a time difference τ:36,86,87
(28) |
(29) |
The resulting tumbling frequencies , normalized with the inverse of the relaxation time, are shown in Fig. 10. As expected, the tumbling frequency grows with the shear rate; however, the resulting growth is much weaker than the dependence of the flexible rings.36 Rigidity slows down the rotational motion of the rings under shear, so that a rigid ring at a specific Weissenberg number tumbles and orients to the flow in a way akin to a flexible ring at some lower Weissenberg number, where tumbling is less frequent. In this way, for the same Weissenberg number, the rigid rings have fewer phases than their flexible counterparts in which they undergo tumbling motions that cause to grow, resulting thereby in a stronger gradient-direction contraction than for flexible cyclic macromolecules.
We finally discuss the effect of the polymers’ presence on the solvent's viscosity, by extracting their intrinsic viscosity , defined as:
(30) |
In eqn (30) above, the polymer stress tensor is calculated in the simulation as a time average:48
(31) |
Our results highlight the interplay between topology, hydrodynamics and thermal fluctuations in dilute polymer solutions. Experimentally, they can in principle be tested in suitably constructed microfluidic devices with single-molecule techniques.39,90,91 Indeed, properties of ring polymers in flows are being elucidated with these techniques33,84,91–93 and even viscoelasticity of supercoiled plasmids has been considered, however only in higher concentration blends so far.5,94 Our results include experimentally accessible chain extent and other characteristics, but to be measured in experiment the DNA plasmids would have to be first sorted according to their supercoiling degree. As our simulations concerned only relatively short rings, the gyration radius depends only weakly on the supercoiling, hence a method other than gel electrophoresis would have to be used. Future directions include the study of the dynamical behaviour in higher concentrations or for topologically linked rigid rings, such as catenanes and the kinetoplast. As a further highly interesting avenue of exploration, particularly in connection to experiments, we see the investigation of longer and/or more flexible rings with torsional stiffness in shear flow. Such rings with high supercoiling degrees display branched structure in equilibrium. The orientation of the branches as well as their restructuring and hydrodynamic swelling connected to the conversion of writhe to twist in shear flow is highly desired, not only due to the biological context, but also as a future perspective tool to tune mechanical properties of polymer solutions using topology in nonequilibrium conditions. The work at hand presents the first step on the way to achieve these goals.
(32) |
(33) |
(34) |
The forces on beads and patches are derived from their gradients with respect to the position where a stands for either a bead or a patch. The force on particle a resulting from the three potentials amounts to:
(35) |
(36) |
(37) |
(38) |
(39) |
(40) |
(41) |
The quaternions form a non-abelian group with respect to their multiplication which is associative. The conjugate of a quaternion is defined as which brings about the quaternion norm and the inverse quaternion q−1 = q*/|q|2, qq−1 = q−1q = 1. A rotation 3 → 3, by an angle α around axis can be expressed by quaternion multiplication. A 3D vector is projected onto the vector quaternion and the rotation is performed with a unit quaternion with , so that97q* = q−1:
(42) |
The rotated vector is read off the resulting vector quaternion . The monomer orientations χi(t) are represented by unit quaternions with . The qi(t) carry the information about a rotation axis and an angle αi corresponding to the orientation of monomer i at time t. The patch positions k = 1, 2, 3 ≡ blue, red, green are recovered by scaling the standard basis vectors to and rotating them with qi(t):
(43) |
In this way, the patch vectors are always guaranteed to both have the correct distance to the bead and be orthogonal to each other so that they span an orthogonal reference system. At the beginning of a simulation run the monomers are given a starting orientation qi(0) that sets the initial patch positions . The second step of the rotational velocity Verlet algorithm (32) comprises a rotation of the orientation qi(t) around axis and by an angle to qi(t + δt). In terms of quaternions it is computed in the following way with :
(44) |
Because of associativity the updated qi(t+δt) imparts the same 3D rotation of a vector to as qi(t) followed by a rotation according to :
(45) |
At time t, the orientation is given by the time-ordered product of all rotations to the left of qi(0):
(46) |
and the step (45) can be repeated t/δt + 1 many times:
(47) |
Inserting and into (45) and (47) makes evident that (43) and (44) reproduce the monomer orientations at all time steps. The quaternions must always have unity norm, i.e., |qi(t)| = 1 so that they represent proper rotations. It holds that , i.e., normalization to unity is guaranteed for all time steps t, if the initial qi(0) have unity norm. The quaternions are normalized every 100th MD step to counteract numerical errors that can possibly occur along the way.
(48) |
Fig. 12 The equilibrium end-to-end correlation function of supercoiled rings of length N = 100 as a function of time starting from the initialization states for + HI, panel (a), and −HI, panel (b). |
Footnotes |
† These authors contributed equally to this work. |
‡ We show in the Appendix that these estimates agree well with the characteristic decay times of the orientational correlation functions of the rings at equilibrium. |
This journal is © The Royal Society of Chemistry 2024 |