Location via proxy:   [ UP ]  
[Report a bug]   [Manage cookies]                
\definecolor

oxfordbluergb0.0, 0.13, 0.28

A first-principles geometric model for dynamics of motor-driven centrosomal asters

Yuan-Nan Young1,3    Vicente Gomez Herrera2    Helena Z. Huan2    Reza Farhadifar3    Michael J. Shelley 111Corresponding author: mshelley@flatironinstitute.org,2,3 1 Department of Mathematical Sciences, New Jersey Institute of Technology, Newark, New Jersey 07102, USA 2 Courant Institute, New York University, NY, NY 10012, USA 3 Center for Computational Biology, Flatiron Institute, NY, NY 10010, USA
Abstract

The centrosomal aster is a mobile and adaptable cellular organelle that exerts and transmits forces necessary for tasks such as nuclear migration and spindle positioning. Recent experimental and theoretical studies of nematode and human cells demonstrate that pulling forces on asters by cortically anchored force generators are dominant during such processes. Here we present a comprehensive investigation of a first-principles model of aster dynamics, the S-model (S for stoichiometry), based solely on such forces. The model evolves the astral centrosome position, a probability field of cell-surface motor occupancy by centrosomal microtubules (under an assumption of stoichiometric binding), and free boundaries of unattached, growing microtubules. We show how cell shape affects the stability of centering of the aster, and its transition to oscillations with increasing motor number. Seeking to understand observations in single-cell nematode embryos, we use highly accurate simulations to examine the nonlinear structures of the bifurcations, and demonstrate the importance of binding domain overlap to interpreting genetic perturbation experiments. We find a generally rich dynamical landscape, dependent upon cell shape, such as internal constant-velocity equatorial orbits of asters that can be seen as traveling wave solutions. Finally, we study the interactions of multiple asters which we demonstrate an effective mutual repulsion due to their competition for surface force generators. We find, amazingly, that centrosomes can relax onto the vertices of platonic and non-platonic solids, very closely mirroring the results of the classical Thomson problem for energy-minimizing configurations of electrons constrained to a sphere and interacting via repulsive Coulomb potentials. Our findings both explain experimental observations, providing insights into the mechanisms governing spindle positioning and cell division dynamics, and show the possibility of new nonlinear phenomena in cell biology.

I Introduction

The centrosome, a micron-scale organelle [1], is the primary microtubule organizing center in animal cells and plays a central role in cellular processes such as division, polarization, and intracellular organization and transport [2]. Microtubules (MTs), stiff polar biopolymers having a persistence length on the order of millimeters [3], nucleate from the centrosome with their plus-ends growing outwards and their minus-ends tethered to the centrosome. A centrosome and its radially oriented MTs form the characteristic centrosomal aster.

An important biophysical aspect of centrosomal asters is their ability to exert and transmit forces. For example, during cell division the mitotic spindle, a bipolar structure primarily composed of transitory MTs and associated proteins, forms near the cell center with a centrosome, and its aster, at each pole (Fig. 1A and Supplementary Video 1). Accurate spindle positioning, largely mediated by the centrosomal asters, is crucial for precise genome and organelle segregation, while errors in centrosome and spindle positioning can prove fatal for cells. Various forces upon centrosomal MTs, including MT polymerization-driven pushing forces against the cell cortex, pulling forces from motor proteins carrying payloads along MTs, pulling forces from cortically anchored motor proteins, and forces from MT friction with the cell wall, have been proposed to drive spindle positioning within cells [4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 20, 21, 22, 23]. While the exact force mechanism behind spindle positioning remains an open question, even for many model organisms, molecular and biophysical perturbation experiments across diverse cell types, including yeast [24], C. elegans [5], and human cells [25], have shown the dominance of pulling forces exerted by cortically anchored motor proteins, such as the minus-end directed motor dynein. A centrosome typically nucleates hundreds of MTs per second, each with a short lifespan of tens of seconds, collectively forming a dynamic aster comprising thousands of MTs. At any given moment a subset of these MTs, typically a few hundred, interfaces with the cortex, giving these MTs the potential to engage with dynein motors anchored there. In this scenario, an impinging MT binds to an anchored dynein motor, which then walks towards the MT’s minus-end and so exerts a pulling force upon it [26]. The dynein motor, together with its anchoring protein complex, is referred to as a cortical force-generators, or CFG. The interaction between CFGs and MTs is transient, perhaps from the motor’s detachment from either cortex or microtubule, or, possibly, the disassembly of the MT itself, but collectively generates the tens of piconewton forces involved in positioning the centrosome, and other structures such as spindles to which it is connected, within the cell [27, 28]. The precise role of these pulling forces in spindle positioning remains the subject of ongoing theoretical and experimental inquiry.

Refer to caption
Figure 1: Mitotic spindle and results from our stoichiometric model. (A) First mitotic spindle in C. elegans embryo. Blue shows the microtubules, pink indicates the chromosomes, and white represents the centrosomes. Scale bar, 10μ10𝜇10\mu10 italic_μm. See also Supplementary Video 1. (B) Equilibrium configuration of two centrosomes (black circles) inside a prolate cell. Color-field represents the probability that a CFG is bound to a MT. (C) Equilibrium configuration of twelve centrosomes that arranged into the vertices of an icosahedron embedded in a sphere.

In prior work, we developed versions of the so-called stoichiometric model, or S-model, for how mobile centrosomal asters interact with CFGs and through them studied aspects of spindle dynamics [23, 29]. The S-model framework is based exclusively on the pulling forces exerted by CFGs upon MTs and, through the transience of their interactions, deals naturally with the cell geometry through which a centrosomal aster moves. It is derived from the fundamental biophysics of MT dynamics and well-established biochemistry of molecular motors. Central is a stoichiometric interaction between CFGs and MTs: a CFG can bind to only one MT at a time. Until detachment, the bound MT experiences a pulling force, which is transmitted to the centrosome. In [23] we developed an S-model with discrete CFGs assuming a quasi-static balance of MT cortical impingement rate and motor-MT detachment. For C. elegans single-cell embryos this model quantitatively explained the dynamics of spindle positioning and elongation, and final length scaling with cell size, while accounting for their variations within species and across nematode species spanning over 100 million years of evolution. In [29], a combination of laser scission of MTs, cytoplasmic flow reconstructions, computational fluid dynamics, and an S-model, were used to demonstrate the predominance of cortical pulling forces in the dynamics of pronuclei and of the spindle in all stages leading up to the first cell division C. elegans embryo. There the S-model was extended to include the dynamics of MT-motor interactions, and the discrete CFGs replaced with a continuous surface distribution of occupancy probability. The S-model consists of an ordinary differential equation (ODE) for centrosome position driven by a surface integral of motor forces weighted by an evolving probability field of MT-motor attachment. Within a spherical cell under symmetry assumptions this S-model was mathematically analyzed to show, among other things, that stable positioning could transition to oscillations via a Hopf bifurcation, at motor densities and at oscillation frequencies consistent with observations of spindles during the metaphase to anaphase transition.

Given the ubiquity and importance of centrosomal asters to cellular dynamics, here we seek to understand this class of models more comprehensively. In § II, we develop an elaborated S-model, which, again, takes the form of an overdamped dynamics for centrosome position driven by a surface integral of coarse-grained MT-oriented pulling forces from CFGs, weighted by the probability for a CFG to be occupied by an MT. The probability evolves as a balance of binding by impinging MTs and unbinding. Unlike our earlier models, here we include the effect of random overlap in CFG binding domains. We show that the S-model has a natural energy-like quantity that evolves through a balance of input power from MT impingement and binding, and dissipation from drag and unbinding. In § III we study the centering of single centrosomes within spheroidal cells; linear analysis for a centrosomal aster predicts stable centering which, with increasing motor density, loses its stability to oscillation via a Hopf bifurcation as suggested by experiment. We find that in the stably centered case, the elaborated S-model accounts for force-displacement experiments using genetic perturbations. Using highly accurate numerical methods we investigate the nonlinear structures of the Hopf bifurcation, seeking and finding supercriticality as is inferred from experiments probing the transition to spindle oscillation. In § IV we show that there is a rich variety of possible dynamics. We find and construct internal equatorial orbits of asters that arise from a symmetry breaking instability. We investigate how multiple centrosomes interact. Within the S-model, centrosomes compete for force-generators, with that competition creating an effective repulsion between them. Consequently, for two centrosomes we can observe relaxation to well-separated positions, as in Fig. 1B, reminiscent of the positioned mitotic spindle, as in Fig. 1A. This competition plays out beautifully when simulating larger numbers where asters move to the vertices of platonic (see Fig. 1C for 12 centrosomes relaxing to an icosahedron) and non-platonic solids in a manner remarkably similar what is found in J.J. Thomson’s classical problem of electrons on a sphere interacting by a Coulomb potential [30].

II A generalized formulation of the stoichiometric model

Refer to caption
Figure 2: Randomly distributed overlapped force generators on the cell surface and schematics of the stoichiometric model. (A) A view of the cell periphery (yellow) with randomly placed CFGs (pink circular disks). (B) Schematics of the S-model. Microtubules (MTs, blue) nucleate from the centrosome (red) with rate γ𝛾\gammaitalic_γ, grow with speed Vgsubscript𝑉𝑔V_{g}italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, and undergo catastrophe with rate λ𝜆\lambdaitalic_λ. The dashed line indicates the microtubule front (𝐒𝐒\mathbf{S}bold_S). Cortical force generators of size rmsubscript𝑟𝑚r_{m}italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT (CFGs, pink) are distributed on the cell surface ΓΓ\Gammaroman_Γ. The interaction of CGFs and MTs is stoichiometric - a CFG can bind only one MT at a time. Once bound, the CFG exerts a pulling force of magnitude f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT along the MT to the centrosome. CFGs detach from MTs with rate κ𝜅\kappaitalic_κ and become unbound.

Building and expanding upon our prior studies [23, 29], we consider the cell periphery to be populated by CFGs in random positions, without considering an exclusion principle (Fig. 2A), and MTs as straight polymers nucleating with spherical uniformity at rate γ𝛾\gammaitalic_γ from a centrosome at location 𝐗(t)𝐗𝑡\mathbf{X}(t)bold_X ( italic_t ); see Fig. 2B. This is consistent with the relative rigidity of MTs and the observation that end-pulling forces create an extensional stress in the attached MTs [31, 32]. With their minus-ends fixed at the centrosome, model MTs grow by polymerization at their plus-ends at a constant speed Vgsubscript𝑉𝑔V_{g}italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, and undergo catastrophe with rate λ𝜆\lambdaitalic_λ; see Fig. 2B.

An MT that reaches the cell surface ΓΓ\Gammaroman_Γ has two possible fates: either it binds to an unoccupied cortical force-generator (CFG) anchored there or undergoes catastrophe and disassemble. We take a stoichiometric interaction between CFGs and MTs: a CFG can only be occupied by one MT at a time. Once a CFG is bound to an MT, it exerts a constant pulling force of magnitude f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT on the centrosome toward the CFG; see Fig. 2B. Bound CFGs will also detach from their MTs with a rate κ𝜅\kappaitalic_κ and become unoccupied. Neglecting inertia, the dynamics of the centrosome and its MT array are governed by a balance between pulling forces from the CFGs and the viscous drag forces from the cytoplasm on the centrosome and its attached MTs.

Consider a centrosome at position 𝐗(t)𝐗𝑡\mathbf{X}(t)bold_X ( italic_t ), hence with velocity 𝐗˙(t)˙𝐗𝑡\dot{\mathbf{X}}(t)over˙ start_ARG bold_X end_ARG ( italic_t ), and at distance D(𝐘,t)=|𝐘𝐗(t)|𝐷𝐘𝑡𝐘𝐗𝑡D(\mathbf{Y},t)=|{\bf Y}-{\bf X}(t)|italic_D ( bold_Y , italic_t ) = | bold_Y - bold_X ( italic_t ) | from points 𝐘Γ𝐘Γ\mathbf{Y}\in\Gammabold_Y ∈ roman_Γ. We define two unit vectors: the outward normal 𝐧^(𝐘)^𝐧𝐘\hat{\mathbf{n}}(\mathbf{Y})over^ start_ARG bold_n end_ARG ( bold_Y ) to ΓΓ\Gammaroman_Γ, and 𝝃^(𝐘;𝐗)=(𝐘𝐗(t))/D(𝐘,t)^𝝃𝐘𝐗𝐘𝐗𝑡𝐷𝐘𝑡\hat{\bm{\xi}}(\mathbf{Y};{\bf X})=({\bf Y}-{\bf X}(t))/D({\bf Y},t)over^ start_ARG bold_italic_ξ end_ARG ( bold_Y ; bold_X ) = ( bold_Y - bold_X ( italic_t ) ) / italic_D ( bold_Y , italic_t ), the direction along which forces are exerted by CFGs.

We assume a total number M𝑀Mitalic_M of CFGs, whose centers are independent and identically distributed with a probability density ρ𝜌\rhoitalic_ρ (i.e. Γρ(𝐘)𝑑A=1subscriptΓ𝜌𝐘differential-d𝐴1\int_{\Gamma}\rho({\bf Y})dA=1∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT italic_ρ ( bold_Y ) italic_d italic_A = 1). Note that independence implies that there is no mutual exclusion between CFGs, hence there can be overlap between the binding domains of different CFGs. Further, we assume that ρ𝜌\rhoitalic_ρ smoothly varies across ΓΓ\Gammaroman_Γ. Note, for a globally uniform distribution of CFGs ρ(𝐘)1/|Γ|𝜌𝐘1Γ\rho(\mathbf{Y})\equiv 1/|\Gamma|italic_ρ ( bold_Y ) ≡ 1 / | roman_Γ | where |Γ|Γ|\Gamma|| roman_Γ | is the surface area of ΓΓ\Gammaroman_Γ.

For sufficiently large M𝑀Mitalic_M, we can apply the law of large numbers, so up to statistical fluctuations, we can express the total pulling force as an integral, and assume the centrosome moves by the overdamped dynamics

η𝐗˙=Mf0ΓP(𝐘,t)𝝃^(𝐘;𝐗)ρ(𝐘)𝑑A+𝐅ext,𝜂˙𝐗𝑀subscript𝑓0subscriptΓ𝑃𝐘𝑡^𝝃𝐘𝐗𝜌𝐘differential-d𝐴subscript𝐅𝑒𝑥𝑡\displaystyle\eta\dot{\bf X}=Mf_{0}\int_{\Gamma}P({\bf Y},t)\hat{\bm{\xi}}({% \bf Y};{\bf X})\,\rho({\bf Y})dA+\mathbf{F}_{ext},italic_η over˙ start_ARG bold_X end_ARG = italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT italic_P ( bold_Y , italic_t ) over^ start_ARG bold_italic_ξ end_ARG ( bold_Y ; bold_X ) italic_ρ ( bold_Y ) italic_d italic_A + bold_F start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT , (1)

where P(𝐘,t)𝑃𝐘𝑡P({\bf Y},t)italic_P ( bold_Y , italic_t ) is the probability that a CFG centered at 𝐘𝐘{\bf Y}bold_Y is occupied by an MT, and η𝜂\etaitalic_η is a drag coefficient. Note that we have allowed for the action of an external force 𝐅extsubscript𝐅𝑒𝑥𝑡\mathbf{F}_{ext}bold_F start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT upon the centrosome so as to mimic particular experiments.

Introducing P(𝐘,t)𝑃𝐘𝑡P({\bf Y},t)italic_P ( bold_Y , italic_t ) instead of the discrete states of a CFG being occupied (bound) or unoccupied (unbound), we have effectively introduced a coarse-graining in time, allowing us to write a forward-Kolmogorov equation. However, to properly do this, we need to do a spatial coarse-graining of the CFGs. In particular, we assume that each CFG can bind over a discal area of radius rmsubscript𝑟𝑚r_{m}italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT (with area am=πrm2subscript𝑎𝑚𝜋superscriptsubscript𝑟𝑚2a_{m}=\pi r_{m}^{2}italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = italic_π italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT) and that P(𝐘,t)𝑃𝐘𝑡P({\bf Y},t)italic_P ( bold_Y , italic_t ) has a slow variation through the surface. Considering stoichiometry of the CFGs, it follows that the evolution equation for the occupancy probability P𝑃Pitalic_P is

P(𝐘,t)t𝑃𝐘𝑡𝑡\displaystyle\frac{\partial P({\bf Y},t)}{\partial t}divide start_ARG ∂ italic_P ( bold_Y , italic_t ) end_ARG start_ARG ∂ italic_t end_ARG =Ω(𝐘,t)(P,𝐘)κP(𝐘,t)withabsentΩ𝐘𝑡𝑃𝐘𝜅𝑃𝐘𝑡with\displaystyle=\Omega(\mathbf{Y},t){\cal I}(P,{\bf Y})-\kappa P({\bf Y},t)~{}% \mbox{with}= roman_Ω ( bold_Y , italic_t ) caligraphic_I ( italic_P , bold_Y ) - italic_κ italic_P ( bold_Y , italic_t ) with (2)
(P,𝐘)𝑃𝐘\displaystyle{\cal I}(P,\mathbf{Y})caligraphic_I ( italic_P , bold_Y ) =1eamMρ(𝐘)(1P(𝐘,t))amMρ(𝐘).absent1superscript𝑒subscript𝑎𝑚𝑀𝜌𝐘1𝑃𝐘𝑡subscript𝑎𝑚𝑀𝜌𝐘\displaystyle=\frac{1-e^{-a_{m}M\rho(\mathbf{Y})(1-P({\bf Y},t))}}{a_{m}M\rho(% \mathbf{Y})}.= divide start_ARG 1 - italic_e start_POSTSUPERSCRIPT - italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_M italic_ρ ( bold_Y ) ( 1 - italic_P ( bold_Y , italic_t ) ) end_POSTSUPERSCRIPT end_ARG start_ARG italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_M italic_ρ ( bold_Y ) end_ARG .

Here Ω(𝐘,t)Ω𝐘𝑡\Omega(\mathbf{Y},t)roman_Ω ( bold_Y , italic_t ) is the MT impingement rate on ΓΓ\Gammaroman_Γ, or the number of unbound MTs per unit time that hit the area covered by a CFG centered at 𝐘𝐘\mathbf{Y}bold_Y, and (P,𝐘)𝑃𝐘{\cal I}(P,{\bf Y})caligraphic_I ( italic_P , bold_Y ) is the probability that an MT succesfully binds to the CFG whose binding domain is centered at 𝐘𝐘{\bf Y}bold_Y (for a complete derivation, see § VI.8).

We can understand (P,𝐘)𝑃𝐘{\cal I}(P,{\bf Y})caligraphic_I ( italic_P , bold_Y ) further by examining its form. First, note that 01010\leq{\cal I}\leq 10 ≤ caligraphic_I ≤ 1 and P<0subscript𝑃0\partial_{P}{\cal I}<0∂ start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT caligraphic_I < 0 for P[0,1]𝑃01P\in[0,1]italic_P ∈ [ 0 , 1 ]. The first bound is trivial, as {\cal I}caligraphic_I is a probability, and the second one is due to the assumption of stoichiometry (the larger the P𝑃Pitalic_P, the less likely a new MT binds to the CFG centered at 𝐘𝐘\bf Ybold_Y). Secondly, amMρ(𝐘)subscript𝑎𝑚𝑀𝜌𝐘a_{m}M\rho(\mathbf{Y})italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_M italic_ρ ( bold_Y ) approximates the number of CFGs covering the point 𝐘𝐘{\bf Y}bold_Y, and thus is a measure of competition between them. For amMρ1much-less-thansubscript𝑎𝑚𝑀𝜌1a_{m}M\rho\ll 1italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_M italic_ρ ≪ 1 (or 1P1much-less-than1𝑃11-P\ll 11 - italic_P ≪ 1), i.e. negligible overlap (or CFGs are close to complete saturation), we have 1P1𝑃{\cal I}\rightarrow 1-Pcaligraphic_I → 1 - italic_P, recovering the independent CFG model, as used in [23, 29]:

P(𝐘,t)t𝑃𝐘𝑡𝑡\displaystyle\frac{\partial P({\bf Y},t)}{\partial t}divide start_ARG ∂ italic_P ( bold_Y , italic_t ) end_ARG start_ARG ∂ italic_t end_ARG =Ω(𝐘,t)(1P(𝐘,t))κP(𝐘,t),absentΩ𝐘𝑡1𝑃𝐘𝑡𝜅𝑃𝐘𝑡\displaystyle=\Omega(\mathbf{Y},t)\left(1-P\left({\bf Y},t\right)\right)-% \kappa P({\bf Y},t),= roman_Ω ( bold_Y , italic_t ) ( 1 - italic_P ( bold_Y , italic_t ) ) - italic_κ italic_P ( bold_Y , italic_t ) ,

where the competition between CFGs becomes irrelevant either due to the low level of overlap or because CFGs are almost always occupied and do not locally compete with each other. For (1P)amMρ1much-greater-than1𝑃subscript𝑎𝑚𝑀𝜌1(1-P)a_{m}M\rho\gg 1( 1 - italic_P ) italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_M italic_ρ ≫ 1, 1amMρ1subscript𝑎𝑚𝑀𝜌{\cal I}\approx\frac{1}{a_{m}M\rho}caligraphic_I ≈ divide start_ARG 1 end_ARG start_ARG italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_M italic_ρ end_ARG, signifying that an MT has an equal chance of binding to any of the CFGs covering the same area, making the competition between CFGs the most relevant. Lastly, the numerator 1eamMρ(𝐘)(1P(𝐘,t))1superscript𝑒subscript𝑎𝑚𝑀𝜌𝐘1𝑃𝐘𝑡1-e^{-a_{m}M\rho(\mathbf{Y})(1-P({\bf Y},t))}1 - italic_e start_POSTSUPERSCRIPT - italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_M italic_ρ ( bold_Y ) ( 1 - italic_P ( bold_Y , italic_t ) ) end_POSTSUPERSCRIPT is the probability that at least one of the CFGs at 𝐘𝐘{\bf Y}bold_Y is unbound (§ VI.8). The term 1amMρ(𝐘)1subscript𝑎𝑚𝑀𝜌𝐘\frac{1}{a_{m}M\rho(\mathbf{Y})}divide start_ARG 1 end_ARG start_ARG italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_M italic_ρ ( bold_Y ) end_ARG comes from the fact that since all CFGs are indistinguishable, the incoming MT will choose one of them randomly.

Following [29] we approximate the impingement rate ΩΩ\Omegaroman_Ω by

Ω(𝐘,t)=ϕ(𝐒)[𝐕S𝐧^]+χ(rmD)(γVgeD/lc).Ω𝐘𝑡italic-ϕ𝐒subscriptdelimited-[]subscript𝐕𝑆^𝐧𝜒subscript𝑟𝑚𝐷𝛾subscript𝑉𝑔superscript𝑒𝐷subscript𝑙𝑐\displaystyle\Omega(\mathbf{Y},t)=\phi({\bf S})\left[{\bf V}_{S}\cdot\hat{\bf n% }\right]_{+}\chi\left(\frac{r_{m}}{D}\right)\left(\frac{\gamma}{V_{g}}e^{-D/l_% {c}}\right).roman_Ω ( bold_Y , italic_t ) = italic_ϕ ( bold_S ) [ bold_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_n end_ARG ] start_POSTSUBSCRIPT + end_POSTSUBSCRIPT italic_χ ( divide start_ARG italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG italic_D end_ARG ) ( divide start_ARG italic_γ end_ARG start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG italic_e start_POSTSUPERSCRIPT - italic_D / italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ) . (3)

In this expression, the first term ϕ(𝐒)italic-ϕ𝐒\phi({\bf S})italic_ϕ ( bold_S ) is the indicator function specifying whether the MT front (Fig. 2B) is in contact with ΓΓ\Gammaroman_Γ (for which ϕ(𝐒)=1italic-ϕ𝐒1\phi({\bf S})=1italic_ϕ ( bold_S ) = 1), or not (for which ϕ(𝐒)=0italic-ϕ𝐒0\phi({\bf S})=0italic_ϕ ( bold_S ) = 0). The second term contains 𝐕S(𝐘,t)=𝐗˙+Vg𝝃^(𝐘;𝐗)subscript𝐕𝑆𝐘𝑡˙𝐗subscript𝑉𝑔^𝝃𝐘𝐗\mathbf{V}_{S}(\mathbf{Y},t)=\dot{\mathbf{X}}+V_{g}\hat{\bm{\xi}}(\mathbf{Y};{% \bf X})bold_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ( bold_Y , italic_t ) = over˙ start_ARG bold_X end_ARG + italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT over^ start_ARG bold_italic_ξ end_ARG ( bold_Y ; bold_X ), the velocity of an MT plus-end with a direction 𝝃^(𝐘;𝐗)^𝝃𝐘𝐗\hat{\bm{\xi}}(\mathbf{Y};{\bf X})over^ start_ARG bold_italic_ξ end_ARG ( bold_Y ; bold_X ), and [𝐕S𝐧^]+subscriptdelimited-[]subscript𝐕𝑆^𝐧\left[{\bf V}_{S}\cdot\hat{\bf n}\right]_{+}[ bold_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_n end_ARG ] start_POSTSUBSCRIPT + end_POSTSUBSCRIPT gives the rectified normal component of that velocity at ΓΓ\Gammaroman_Γ; If 𝐕S𝐧^<0subscript𝐕𝑆^𝐧0{\bf V}_{S}\cdot\hat{\bf n}<0bold_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_n end_ARG < 0 the centrosome is moving away from the surface faster than MTs grow, and thus the impingement rate is zero. The third term χ(rm/D)=12(111+(rm/D)2)𝜒subscript𝑟𝑚𝐷12111superscriptsubscript𝑟𝑚𝐷2\chi\left(r_{m}/D\right)=\frac{1}{2}\left(1-\frac{1}{\sqrt{1+(r_{m}/D)^{2}}}\right)italic_χ ( italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT / italic_D ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 1 - divide start_ARG 1 end_ARG start_ARG square-root start_ARG 1 + ( italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT / italic_D ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG ) estimates the fraction of centrosomally-nucleated, unbound MTs at ΓΓ\Gammaroman_Γ available for binding to a CFG. Finally, γVgeD/lc𝛾subscript𝑉𝑔superscript𝑒𝐷subscript𝑙𝑐\frac{\gamma}{V_{g}}e^{-D/l_{c}}divide start_ARG italic_γ end_ARG start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG italic_e start_POSTSUPERSCRIPT - italic_D / italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUPERSCRIPT is the equilibrium astral MT number per unit length reaching ΓΓ\Gammaroman_Γ from the centrosome (at distance D𝐷Ditalic_D), with lc=Vg/λsubscript𝑙𝑐subscript𝑉𝑔𝜆l_{c}=V_{g}/\lambdaitalic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT / italic_λ the characteristic length of MTs undergoing dynamic instability. We note that for time t>W/Vg𝑡𝑊subscript𝑉𝑔t>W/V_{g}italic_t > italic_W / italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT it is appropriate to use the steady state distribution because of linear hyperbolic nature of the associated Fokker-Plank equation for the MT length distribution [29].

Next, we consider the dynamics of the MT front (surface 𝐒𝐒{\bf S}bold_S in Fig. 2B), which incorporates the finite propagation speed of MT plus-ends and accounts for the detachment of the MT front due to the centrosome moving away from the cell boundary faster than MT growth. The full front is described by the equation

d𝐒dt𝑑𝐒𝑑𝑡\displaystyle\frac{d{\bf S}}{dt}divide start_ARG italic_d bold_S end_ARG start_ARG italic_d italic_t end_ARG =ϕ(𝐒)[𝐕S𝐧^]+𝐕S+(1ϕ(𝐒))𝐕S,absentitalic-ϕ𝐒subscriptdelimited-[]subscript𝐕𝑆^𝐧subscript𝐕𝑆1italic-ϕ𝐒subscript𝐕𝑆\displaystyle=\phi({\bf S})\left[-{\bf V}_{S}\cdot\hat{\bf n}\right]_{+}{\bf V% }_{S}+\left(1-\phi({\bf S})\right){\bf V}_{S}\,,= italic_ϕ ( bold_S ) [ - bold_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_n end_ARG ] start_POSTSUBSCRIPT + end_POSTSUBSCRIPT bold_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT + ( 1 - italic_ϕ ( bold_S ) ) bold_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT , (4)

Note that when the front is not located in the cell periphery, it translates with the centrosome expanding at velocity Vgsubscript𝑉𝑔V_{g}italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, but when a point on the MT front is located on the cell periphery, it either stays on the periphery or moves with the velocity 𝐕Ssubscript𝐕𝑆{\bf V}_{S}bold_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT inwards and away from the cell boundary.

The model, Eqs. (1-4), has seven parameters (see Table 1). We take the cell scale W=|Γ|/4π𝑊Γ4𝜋W=\sqrt{|\Gamma|/4\pi}italic_W = square-root start_ARG | roman_Γ | / 4 italic_π end_ARG as a characteristic length scale, and τ=W/Vg𝜏𝑊subscript𝑉𝑔\tau=W/V_{g}italic_τ = italic_W / italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT (the time scale for an MT to grow from cell center to cell periphery) as a characteristic time scale. Other faster time scales are the inverse nucleation rate 1/γ1𝛾1/\gamma1 / italic_γ (fastest) and the inverse detachment rate 1/κ1𝜅1/\kappa1 / italic_κ (next fastest). The smallest length scale is rmsubscript𝑟𝑚r_{m}italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT. From these we obtain five dimensionless parameters: dimensionless nucleation rate γ¯=τγ¯𝛾𝜏𝛾\bar{\gamma}=\tau\gammaover¯ start_ARG italic_γ end_ARG = italic_τ italic_γ, dimensionless detachment rate κ¯=τκ¯𝜅𝜏𝜅\bar{\kappa}=\tau\kappaover¯ start_ARG italic_κ end_ARG = italic_τ italic_κ, dimensionless MT length l¯c=lc/Wsubscript¯𝑙𝑐subscript𝑙𝑐𝑊\bar{l}_{c}=l_{c}/Wover¯ start_ARG italic_l end_ARG start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_W, dimensionless CFG size r¯m=rm/Wsubscript¯𝑟𝑚subscript𝑟𝑚𝑊\bar{r}_{m}=r_{m}/Wover¯ start_ARG italic_r end_ARG start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT = italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT / italic_W, and dimensionless force f¯0=f0/ηVgsubscript¯𝑓0subscript𝑓0𝜂subscript𝑉𝑔\bar{f}_{0}=f_{0}/\eta V_{g}over¯ start_ARG italic_f end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_η italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT.

Comments:

i. The spatial coupling of P(𝐘,t)𝑃𝐘𝑡P(\mathbf{Y},t)italic_P ( bold_Y , italic_t ) occurs only through the integral coupling provided by Eq. (1), and not through the explicit P𝑃Pitalic_P dynamics, Eq. (2), which is entirely local in 𝐘𝐘\mathbf{Y}bold_Y. Eqs. (1,2) are a countably infinite set of ODEs when P𝑃Pitalic_P is expanded in a countable and complete set of surface basis functions (such as spherical harmonics).

ii. The probability (P)𝑃{\cal I}(P)caligraphic_I ( italic_P ) can take different forms, depending upon modeling assumptions. As discussed above, in the limit amM/|Γ|1much-less-thansubscript𝑎𝑚𝑀Γ1a_{m}M/|\Gamma|\ll 1italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_M / | roman_Γ | ≪ 1, where the overlap between CFGs is negligible, (P)𝑃{\cal I}(P)caligraphic_I ( italic_P ) in Eq. (2) simplifies to =1P1𝑃{\cal I}=1-Pcaligraphic_I = 1 - italic_P, and the occupancy probability satisfies the independent CFG model. This was the form used in [29], and earlier in discrete form by [23]. In recent work in human spindles, we consider localized clusters of N𝑁Nitalic_N CFGs that are independent between each other, which yields an interaction in the form =1PNN1superscript𝑃𝑁𝑁{\cal I}=\frac{1-P^{N}}{N}caligraphic_I = divide start_ARG 1 - italic_P start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT end_ARG start_ARG italic_N end_ARG.

iii. Formally, one can write the RHS of Eq. (1) as the gradient of an energy

η𝐗˙𝜂˙𝐗\displaystyle\eta\dot{\bf X}italic_η over˙ start_ARG bold_X end_ARG =𝐗,absentsubscript𝐗\displaystyle=-\nabla_{{\bf X}}{\cal E},= - ∇ start_POSTSUBSCRIPT bold_X end_POSTSUBSCRIPT caligraphic_E , (5)

where the energy {\cal E}caligraphic_E is defined as

[P;𝐗]𝑃𝐗\displaystyle{\cal E}[P;\mathbf{X}]caligraphic_E [ italic_P ; bold_X ] =Mf0ΓP(𝐘,t)D(𝐘,t)ρ(𝐘)𝑑A0,absent𝑀subscript𝑓0subscriptΓ𝑃𝐘𝑡𝐷𝐘𝑡𝜌𝐘differential-d𝐴0\displaystyle=Mf_{0}\int_{\Gamma}P({\bf Y},t)D({\bf Y},t)\rho({\bf Y})dA\geq 0,= italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT italic_P ( bold_Y , italic_t ) italic_D ( bold_Y , italic_t ) italic_ρ ( bold_Y ) italic_d italic_A ≥ 0 , (6)

recalling that D(𝐘,t)=|𝐘𝐗(t)|𝐷𝐘𝑡𝐘𝐗𝑡D(\mathbf{Y},t)=\left|\mathbf{Y}-\mathbf{X}(t)\right|italic_D ( bold_Y , italic_t ) = | bold_Y - bold_X ( italic_t ) |. This energy achieves its minimum of zero only for P0𝑃0P\equiv 0italic_P ≡ 0, that is when there are no bound microtubules (and no velocity). Were P𝑃Pitalic_P independent of time, which it is typically not, the energy would change only with the centrosome position and so {\cal E}caligraphic_E would decay in time due to Eq. (5). Instead, using Eqs. (1,2), {\cal E}caligraphic_E evolves as

˙˙\displaystyle\dot{\cal E}over˙ start_ARG caligraphic_E end_ARG =Mf0ΓΩ(𝐘,t)(P)D(𝐘,t)ρ(𝐘)𝑑A(κ+η|𝐗˙|2)=𝒫𝒟,absent𝑀subscript𝑓0subscriptΓΩ𝐘𝑡𝑃𝐷𝐘𝑡𝜌𝐘differential-d𝐴𝜅𝜂superscript˙𝐗2𝒫𝒟\displaystyle=Mf_{0}\int_{\Gamma}\Omega({\bf Y},t){\cal I}(P)D({\bf Y},t)\rho(% {\bf Y})dA-\left(\kappa{\cal E}+\eta\left|\dot{\bf X}\right|^{2}\right)=\cal{P% }-\cal{D},= italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT roman_Ω ( bold_Y , italic_t ) caligraphic_I ( italic_P ) italic_D ( bold_Y , italic_t ) italic_ρ ( bold_Y ) italic_d italic_A - ( italic_κ caligraphic_E + italic_η | over˙ start_ARG bold_X end_ARG | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) = caligraphic_P - caligraphic_D , (7)

where 𝒫0𝒫0{\cal P}\geq 0caligraphic_P ≥ 0 is the input power due to microtubule impingement and binding, and 𝒟0𝒟0{\cal D}\geq 0caligraphic_D ≥ 0 are the dissipations arising from microtubule detachment from CFGs and the viscous drag of aster motion. Hence, if {\cal E}caligraphic_E were in a nonzero steady-state this would reflect a balance of positive input power and positive dissipation.

iv. The S-model generalizes naturally to the dynamics of N𝑁Nitalic_N centrosomes, with each (ithsuperscript𝑖𝑡i^{th}italic_i start_POSTSUPERSCRIPT italic_t italic_h end_POSTSUPERSCRIPT) centrosome at its position 𝐗i(t)superscript𝐗𝑖𝑡{\bf X}^{i}(t)bold_X start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( italic_t ) having its surface probability Pi(𝐘,t)superscript𝑃𝑖𝐘𝑡P^{i}({\bf Y},t)italic_P start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( bold_Y , italic_t ) of CFG binding. In this generalization, all centrosomal asters compete for the CFGs, which is a fixed resource, through their impingement rates Ωi(𝐘)superscriptΩ𝑖𝐘\Omega^{i}({\bf Y})roman_Ω start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( bold_Y ). This competition is found in the dynamics of Pi(𝐘,t)superscript𝑃𝑖𝐘𝑡P^{i}({\bf Y},t)italic_P start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( bold_Y , italic_t ):

Pi(𝐘,t)t=Ωi(𝐘,t)(P)κPi(𝐘,t),superscript𝑃𝑖𝐘𝑡𝑡superscriptΩ𝑖𝐘𝑡𝑃𝜅superscript𝑃𝑖𝐘𝑡\frac{\partial P^{i}({\bf Y},t)}{\partial t}=\Omega^{i}(\mathbf{Y},t){\cal I}(% P)-\kappa P^{i}({\bf Y},t),divide start_ARG ∂ italic_P start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( bold_Y , italic_t ) end_ARG start_ARG ∂ italic_t end_ARG = roman_Ω start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( bold_Y , italic_t ) caligraphic_I ( italic_P ) - italic_κ italic_P start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( bold_Y , italic_t ) , (8)

where P=i=1NPi𝑃subscriptsuperscript𝑁𝑖1superscript𝑃𝑖P=\sum\limits^{N}\limits_{i=1}P^{i}italic_P = ∑ start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT italic_P start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT is the total CFG occupancy probability. The total probability P𝑃Pitalic_P then satisfies

P(𝐘,t)t𝑃𝐘𝑡𝑡\displaystyle\frac{\partial P({\bf Y},t)}{\partial t}divide start_ARG ∂ italic_P ( bold_Y , italic_t ) end_ARG start_ARG ∂ italic_t end_ARG =(i=1NΩi(𝐘,t))(P)κP(𝐘,t).absentsubscriptsuperscript𝑁𝑖1superscriptΩ𝑖𝐘𝑡𝑃𝜅𝑃𝐘𝑡\displaystyle=\left(\sum\limits^{N}\limits_{i=1}\Omega^{i}(\mathbf{Y},t)\right% ){\cal I}\left(P\right)-\kappa P({\bf Y},t).= ( ∑ start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT roman_Ω start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( bold_Y , italic_t ) ) caligraphic_I ( italic_P ) - italic_κ italic_P ( bold_Y , italic_t ) . (9)

The earlier S-model of [23], using N=2𝑁2N=2italic_N = 2, was used there to study the effects of motor competition on the dynamics and steady-states of two asters, as a way of understanding spindle length selection. Likewise, there is a multi-centrosome analog to Eqs. (6,7) for a total energy =i=1Nisubscriptsuperscript𝑁𝑖1superscript𝑖{\cal E}=\sum\limits^{N}\limits_{i=1}{\cal E}^{i}caligraphic_E = ∑ start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT caligraphic_E start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT.

II.1 Simulation Methods

To numerically evolve 𝐗𝐗{\bf X}bold_X and P𝑃Pitalic_P via Eqs. (1) and (2), respectively, we first compute the surface integral of P𝑃Pitalic_P against 𝝃^^𝝃\hat{\bm{\xi}}over^ start_ARG bold_italic_ξ end_ARG in Eq. (1), given 𝐗𝐗{\bf X}bold_X and P𝑃Pitalic_P. Two different quadrature schemes are used to evaluate the surface integral: For the simulations in spheres examining oscillations near the Hopf bifurcation (§ III.4), P𝑃Pitalic_P is expanded in Legendre polynomials (§ VI.3) and a high-order Gauss-Legendre quadrature is used to evaluate the integral. Otherwise, we map the cell surface to the unit sphere and use a more general patch-based high-order quadrature (§ VI.5). The centrosome position is updated using the explicit Euler method with a time-step ΔtΔ𝑡\Delta troman_Δ italic_t which also yields 𝐗˙˙𝐗\dot{\bf X}over˙ start_ARG bold_X end_ARG at the current time-step. Using this centrosome velocity and accounting for MT growth we then update the position of the microtubule front 𝐒𝐒\bf{S}bold_S position via an Euler step, whether 𝐒𝐒\mathbf{S}bold_S was in contact with ΓΓ\Gammaroman_Γ or not. There are two cases. If the updated 𝐒𝐒\mathbf{S}bold_S lies within the cell, i.e. remains or becomes detached, then the impingement rate imposed upon the line-of-sight CFGs (i.e. from the centrosome) is zero. If the front has moved outside of the cell then the front is projected back to ΓΓ\Gammaroman_Γ and ΩΩ\Omegaroman_Ω is calculated for updating P𝑃Pitalic_P in those regions. For stability, we use a mixed explicit/implicit Euler scheme. The term (P)𝑃{\cal I}(P)caligraphic_I ( italic_P ) is treated explicitly, while the damping term κP𝜅𝑃-\kappa P- italic_κ italic_P is treated implicitly. This yields an explicit expression for P𝑃Pitalic_P at the next time step (Eq. (26) in § VI.2). The local error in time-stepping is found to be quadratic in ΔtΔ𝑡\Delta troman_Δ italic_t, and the global error scales linearly with ΔtΔ𝑡\Delta troman_Δ italic_t, as expected.

III Bifurcations and their nonlinear structures

Refer to caption
Figure 3: Linear analysis of the S-model and on-axis centrosome dynamics in the sphere. (A) Linear solutions of the eigenvalues as a function of the bifurcation parameter M𝑀Mitalic_M. Values of M𝑀Mitalic_M where a change in the centrosome dynamics occurs, are indicated. For M<M1=166𝑀subscript𝑀1166M<M_{1}=166italic_M < italic_M start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = 166, the centrosome stably positions at the cell center. For M1<M<MH=526subscript𝑀1𝑀subscript𝑀𝐻526M_{1}<M<M_{H}=526italic_M start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT < italic_M < italic_M start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT = 526, the initially displaced centrosome moves to the center with damped oscillation. For MH<M<M2=883subscript𝑀𝐻𝑀subscript𝑀2883M_{H}<M<M_{2}=883italic_M start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT < italic_M < italic_M start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 883, the centrosome oscillates around the center with a growing amplitude that plateaus. For M>M2𝑀subscript𝑀2M>M_{2}italic_M > italic_M start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, the center is unstable and the centrosome decenters without oscillation. For M>Mc𝑀subscript𝑀𝑐M>M_{c}italic_M > italic_M start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, the coefficient of the force-displacement changes sign. (B) Examples of centrosome trajectories for various values of M𝑀Mitalic_M as a function of time are shown.

We first examine the linear behavior of the generalized S-model under small perturbations. For ease of calculation we assume that cell shapes are convex, axisymmetric about the z𝑧zitalic_z-axis, and symmetric across the xy𝑥𝑦xyitalic_x italic_y-plane. Thus the origin, or the cell center X0subscript𝑋0X_{0}italic_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, is naturally a fixed point of centrosome position. It has an associated steady-state occupancy probability P=P0(𝐘)𝑃subscript𝑃0𝐘P=P_{0}(\mathbf{Y})italic_P = italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) satisfying

ΓP0(𝐘)𝐘^ρ(𝐘)𝑑AY=𝟎,andP0(𝐘)=Ω0(𝐘)(P0)/κformulae-sequencesubscriptΓsubscript𝑃0𝐘^𝐘𝜌𝐘differential-dsubscript𝐴𝑌0andsubscript𝑃0𝐘subscriptΩ0𝐘subscript𝑃0𝜅\int_{\Gamma}P_{0}(\mathbf{Y})\hat{\mathbf{Y}}\,\rho(\mathbf{Y})dA_{Y}=\mathbf% {0},~{}~{}\mbox{and}~{}~{}P_{0}({\bf Y})=\Omega_{0}({\bf Y}){\cal I}\left(P_{0% }\right)/\kappa∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) over^ start_ARG bold_Y end_ARG italic_ρ ( bold_Y ) italic_d italic_A start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT = bold_0 , and italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) = roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) caligraphic_I ( italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) / italic_κ (10)

where

Ω0(𝐘)=𝐘^𝐧^(𝐘)χ(rm|𝐘|)e|𝐘|lcγsubscriptΩ0𝐘^𝐘^𝐧𝐘𝜒subscript𝑟𝑚𝐘superscript𝑒𝐘subscript𝑙𝑐𝛾\Omega_{0}({\mathbf{Y}})=\hat{\mathbf{Y}}\cdot\hat{\mathbf{n}}({\bf Y})~{}\chi% \left(\frac{r_{m}}{|\mathbf{Y}|}\right)e^{-\frac{|\mathbf{Y}|}{l_{c}}}\gammaroman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) = over^ start_ARG bold_Y end_ARG ⋅ over^ start_ARG bold_n end_ARG ( bold_Y ) italic_χ ( divide start_ARG italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG | bold_Y | end_ARG ) italic_e start_POSTSUPERSCRIPT - divide start_ARG | bold_Y | end_ARG start_ARG italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG end_POSTSUPERSCRIPT italic_γ (11)

and 𝐘^=𝐘/|𝐘|^𝐘𝐘𝐘\hat{\mathbf{Y}}=\mathbf{Y}/|\mathbf{Y}|over^ start_ARG bold_Y end_ARG = bold_Y / | bold_Y |. Note that, under these conditions, cell convexity implies that the MT front is in contact with the cell surface at all points (ϕ(𝐒)1italic-ϕ𝐒1\phi(\mathbf{S})\equiv 1italic_ϕ ( bold_S ) ≡ 1). Also, for a general cell shape the nonlinear equation for P0subscript𝑃0P_{0}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT must typically be solved numerically.

Consider perturbations around the fixed point, 𝐗(t)=ε𝐱(t)𝐗𝑡𝜀𝐱𝑡\mathbf{X}(t)=\varepsilon\mathbf{x}(t)bold_X ( italic_t ) = italic_ε bold_x ( italic_t ) (𝐗˙=ε𝐱˙˙𝐗𝜀˙𝐱\dot{\mathbf{X}}=\varepsilon\dot{\mathbf{x}}over˙ start_ARG bold_X end_ARG = italic_ε over˙ start_ARG bold_x end_ARG) and P(𝐘,t)=P0(𝐘)+εP1(𝐘,t)𝑃𝐘𝑡subscript𝑃0𝐘𝜀subscript𝑃1𝐘𝑡P({\bf Y},t)=P_{0}({\bf Y})+\varepsilon P_{1}({\bf Y},t)italic_P ( bold_Y , italic_t ) = italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) + italic_ε italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_Y , italic_t ), likewise with a small external force 𝐅ext=ε𝐟extsubscript𝐅𝑒𝑥𝑡𝜀subscript𝐟𝑒𝑥𝑡\mathbf{F}_{ext}=\varepsilon\mathbf{f}_{ext}bold_F start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT = italic_ε bold_f start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT, where ε1much-less-than𝜀1\varepsilon\ll 1italic_ε ≪ 1. We can also assume |𝐗˙|Vgmuch-less-than˙𝐗subscript𝑉𝑔|\dot{\mathbf{X}}|\ll V_{g}| over˙ start_ARG bold_X end_ARG | ≪ italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, ensuring continuous contact between the MT front and the cell surface. Therefore, we can disregard the MT front dynamics in Eq. (4). Expanding in ε𝜀\varepsilonitalic_ε and truncating at linear order, from Eqs. (1,2) we find the linearized equations of motion to be

η~𝐱˙(t)=ΓP0(𝐘)|𝐘|(𝐈𝐘^𝐘^)ρ(𝐘)𝑑AY𝐱(t)+ΓP1(𝐘,t)𝐘^ρ(𝐘)𝑑AY+𝐟~ext.~𝜂˙𝐱𝑡subscriptΓsubscript𝑃0𝐘𝐘𝐈^𝐘^𝐘𝜌𝐘differential-dsubscript𝐴𝑌𝐱𝑡subscriptΓsubscript𝑃1𝐘𝑡^𝐘𝜌𝐘differential-dsubscript𝐴𝑌subscript~𝐟𝑒𝑥𝑡\displaystyle\tilde{\eta}\dot{\bf x}(t)=-\int_{\Gamma}\frac{P_{0}({\bf Y})}{|% \mathbf{Y}|}\left({\bf I}-\hat{\mathbf{Y}}\hat{\mathbf{Y}}\right)\rho({\bf Y})% dA_{Y}\cdot{\bf x}(t)+\int_{\Gamma}P_{1}({\bf Y},t)\hat{\mathbf{Y}}\rho({\bf Y% })dA_{Y}+\tilde{\mathbf{f}}_{ext}.over~ start_ARG italic_η end_ARG over˙ start_ARG bold_x end_ARG ( italic_t ) = - ∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT divide start_ARG italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) end_ARG start_ARG | bold_Y | end_ARG ( bold_I - over^ start_ARG bold_Y end_ARG over^ start_ARG bold_Y end_ARG ) italic_ρ ( bold_Y ) italic_d italic_A start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ⋅ bold_x ( italic_t ) + ∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_Y , italic_t ) over^ start_ARG bold_Y end_ARG italic_ρ ( bold_Y ) italic_d italic_A start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT + over~ start_ARG bold_f end_ARG start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT . (12)
P1t(𝐘,t)=κP0(𝐘)𝐧^(𝐘)(𝐱˙(t)Vg+𝐀1(𝐘)𝐱(t))Ω~0(𝐘)P1(𝐘,t),subscript𝑃1𝑡𝐘𝑡𝜅subscript𝑃0𝐘^𝐧𝐘˙𝐱𝑡subscript𝑉𝑔subscript𝐀1𝐘𝐱𝑡subscript~Ω0𝐘subscript𝑃1𝐘𝑡\displaystyle\frac{\partial P_{1}}{\partial t}({\bf Y},t)=\kappa P_{0}({\bf Y}% )\,\hat{\mathbf{n}}({\bf Y})\cdot\left(\frac{\dot{\bf x}(t)}{V_{g}}+{\bf A}_{1% }({\bf Y})\cdot{\bf x}(t)\right)-\tilde{\Omega}_{0}({\bf Y})P_{1}({\bf Y},t),divide start_ARG ∂ italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_t end_ARG ( bold_Y , italic_t ) = italic_κ italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) over^ start_ARG bold_n end_ARG ( bold_Y ) ⋅ ( divide start_ARG over˙ start_ARG bold_x end_ARG ( italic_t ) end_ARG start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG + bold_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_Y ) ⋅ bold_x ( italic_t ) ) - over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_Y , italic_t ) , (13)

where η~=ηMf0~𝜂𝜂𝑀subscript𝑓0\tilde{\eta}=\frac{\eta}{Mf_{0}}over~ start_ARG italic_η end_ARG = divide start_ARG italic_η end_ARG start_ARG italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG and 𝐟~ext=𝐟extMf0subscript~𝐟𝑒𝑥𝑡subscript𝐟𝑒𝑥𝑡𝑀subscript𝑓0\tilde{\mathbf{f}}_{ext}=\frac{\mathbf{f}_{ext}}{Mf_{0}}over~ start_ARG bold_f end_ARG start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT = divide start_ARG bold_f start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT end_ARG start_ARG italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG, and with

𝐀1(𝐘)subscript𝐀1𝐘\displaystyle{\bf A}_{1}({\mathbf{Y}})bold_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_Y ) =1|𝐘|[𝐈(1+|𝐘|lc+rm|𝐘|χ(rm|𝐘|)χ(rm|𝐘|))𝐘^𝐘^],absent1𝐘delimited-[]𝐈1𝐘subscript𝑙𝑐subscript𝑟𝑚𝐘superscript𝜒subscript𝑟𝑚𝐘𝜒subscript𝑟𝑚𝐘^𝐘^𝐘\displaystyle=-\frac{1}{\lvert\mathbf{Y}\rvert}\left[\mathbf{I}-\left(1+\frac{% |\mathbf{Y}|}{l_{c}}+\frac{r_{m}}{|\mathbf{Y}|}\frac{\chi^{\prime}\left(\frac{% r_{m}}{|\mathbf{Y}|}\right)}{\chi\left(\frac{r_{m}}{|\mathbf{Y}|}\right)}% \right)\hat{\mathbf{Y}}\hat{\mathbf{Y}}\right],= - divide start_ARG 1 end_ARG start_ARG | bold_Y | end_ARG [ bold_I - ( 1 + divide start_ARG | bold_Y | end_ARG start_ARG italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG + divide start_ARG italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG | bold_Y | end_ARG divide start_ARG italic_χ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( divide start_ARG italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG | bold_Y | end_ARG ) end_ARG start_ARG italic_χ ( divide start_ARG italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG | bold_Y | end_ARG ) end_ARG ) over^ start_ARG bold_Y end_ARG over^ start_ARG bold_Y end_ARG ] , (14)
Ω~0(𝐘)subscript~Ω0𝐘\displaystyle\tilde{\Omega}_{0}({\bf Y})over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) =κ+Ω0(𝐘)eamMρ(𝐘)(1P0(𝐘))>0.absent𝜅subscriptΩ0𝐘superscript𝑒subscript𝑎𝑚𝑀𝜌𝐘1subscript𝑃0𝐘0\displaystyle=\kappa+\Omega_{0}({\bf Y})e^{-a_{m}M\rho\left({\bf Y}\right)(1-P% _{0}({\bf Y}))}>0.= italic_κ + roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) italic_e start_POSTSUPERSCRIPT - italic_a start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT italic_M italic_ρ ( bold_Y ) ( 1 - italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) ) end_POSTSUPERSCRIPT > 0 . (15)

The matrix 𝐀1subscript𝐀1{\bf A}_{1}bold_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is symmetric and nonsingular.

We first perform linear stability analysis of Eqs. (12,13) in the absence of an external force (𝐟ext=𝟎subscript𝐟𝑒𝑥𝑡0{\mathbf{f}}_{ext}={\mathbf{0}}bold_f start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT = bold_0) and, second, compute the linear force-displacement relation for a stationary centrosome (𝐱˙=𝟎˙𝐱0\dot{{\mathbf{x}}}={\mathbf{0}}over˙ start_ARG bold_x end_ARG = bold_0). Analytical results are derived for spherical geometry, and semi-analytic extensions are obtained for spheroidal geometries. For simplicity, we now assume the CFG distribution is uniform (ρ1/|Γ|𝜌1Γ\rho\equiv 1/|\Gamma|italic_ρ ≡ 1 / | roman_Γ |).

III.1 Linearized dynamics in a spherical cell

Consider a spherical cell of radius W𝑊Witalic_W, for which the equations above now simplifies drastically. In this case 𝐘^=𝐧^^𝐘^𝐧\hat{\mathbf{Y}}=\hat{\mathbf{n}}over^ start_ARG bold_Y end_ARG = over^ start_ARG bold_n end_ARG, |𝐘|=W𝐘𝑊|\mathbf{Y}|=W| bold_Y | = italic_W, and Ω0subscriptΩ0\Omega_{0}roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, Ω~0subscript~Ω0\tilde{\Omega}_{0}over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, and P0subscript𝑃0P_{0}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT are all constants. In brief sketch, we use the fact that 𝐧^𝐀1𝐘^conditional^𝐧subscript𝐀1^𝐘\hat{\bf n}\cdot\mathbf{A}_{1}\parallel\hat{\mathbf{Y}}over^ start_ARG bold_n end_ARG ⋅ bold_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ∥ over^ start_ARG bold_Y end_ARG and the identity Γ𝐘^𝐘^𝑑AY=|Γ|3𝐈subscriptΓ^𝐘^𝐘differential-dsubscript𝐴𝑌Γ3𝐈\int_{\Gamma}\hat{\mathbf{Y}}\hat{\mathbf{Y}}dA_{Y}=\frac{|\Gamma|}{3}{\bf I}∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT over^ start_ARG bold_Y end_ARG over^ start_ARG bold_Y end_ARG italic_d italic_A start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT = divide start_ARG | roman_Γ | end_ARG start_ARG 3 end_ARG bold_I to show that Eqs. (12,13) take the form

𝐱˙=a𝐱+bΓP1𝐘^𝑑SY+𝐠extandP1t=𝐕(t)𝐘^Ω~0P1,˙𝐱𝑎𝐱𝑏subscriptΓsubscript𝑃1^𝐘differential-dsubscript𝑆𝑌subscript𝐠𝑒𝑥𝑡andsubscript𝑃1𝑡𝐕𝑡^𝐘subscript~Ω0subscript𝑃1\displaystyle\dot{\mathbf{x}}=-a\mathbf{x}+b\int_{\Gamma}P_{1}\hat{\mathbf{Y}}% dS_{Y}+\mathbf{g}_{ext}~{}~{}\mbox{and}~{}~{}P_{1t}=\mathbf{V}(t)\cdot\hat{% \mathbf{Y}}-\tilde{\Omega}_{0}P_{1},over˙ start_ARG bold_x end_ARG = - italic_a bold_x + italic_b ∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT over^ start_ARG bold_Y end_ARG italic_d italic_S start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT + bold_g start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT and italic_P start_POSTSUBSCRIPT 1 italic_t end_POSTSUBSCRIPT = bold_V ( italic_t ) ⋅ over^ start_ARG bold_Y end_ARG - over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ,

where 𝐕=α𝐱+β𝐱˙𝐕𝛼𝐱𝛽˙𝐱\mathbf{V}=\alpha\mathbf{x}+\beta\dot{\mathbf{x}}bold_V = italic_α bold_x + italic_β over˙ start_ARG bold_x end_ARG, and a𝑎aitalic_a, b𝑏bitalic_b, α𝛼\alphaitalic_α, and β𝛽\betaitalic_β are all positive constants and 𝐠ext=𝐟~ext/η~subscript𝐠𝑒𝑥𝑡subscript~𝐟𝑒𝑥𝑡~𝜂{\bf g}_{ext}={\tilde{\bf f}}_{ext}/{\tilde{\eta}}bold_g start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT = over~ start_ARG bold_f end_ARG start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT / over~ start_ARG italic_η end_ARG. We note that ΓP1𝐘^𝑑SYsubscriptΓsubscript𝑃1^𝐘differential-dsubscript𝑆𝑌\int_{\Gamma}P_{1}\hat{\mathbf{Y}}dS_{Y}∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT over^ start_ARG bold_Y end_ARG italic_d italic_S start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT is the projection of P1subscript𝑃1P_{1}italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT against the three spherical harmonics (the components of 𝐘^^𝐘\hat{\bf Y}over^ start_ARG bold_Y end_ARG) with l=1𝑙1l=1italic_l = 1 polar index. A linear combination of these three l=1𝑙1l=1italic_l = 1 harmonics also emerges in the forcing term, 𝐕(t)𝐘^𝐕𝑡^𝐘\mathbf{V}(t)\cdot\hat{\mathbf{Y}}bold_V ( italic_t ) ⋅ over^ start_ARG bold_Y end_ARG, in the evolution equation for P1subscript𝑃1P_{1}italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. Consequently this system can be decomposed by setting P1=𝐄𝐘^+Q(𝐘,t)subscript𝑃1𝐄^𝐘𝑄𝐘𝑡P_{1}=\mathbf{E}\cdot\hat{\mathbf{Y}}+Q(\mathbf{Y},t)italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = bold_E ⋅ over^ start_ARG bold_Y end_ARG + italic_Q ( bold_Y , italic_t ) where Q𝑄Qitalic_Q is the projection of P1subscript𝑃1P_{1}italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT onto the spherical harmonics with l1𝑙1l\neq 1italic_l ≠ 1, that is, those orthogonal to 𝐘^^𝐘\hat{\mathbf{Y}}over^ start_ARG bold_Y end_ARG. We then find

𝐱˙=a𝐱+b|Γ|3𝐄+𝐠ext,𝐄˙=(α𝐱+β𝐱˙)Ω~0𝐄,andQt=Ω~0Qformulae-sequence˙𝐱𝑎𝐱𝑏Γ3𝐄subscript𝐠𝑒𝑥𝑡formulae-sequence˙𝐄𝛼𝐱𝛽˙𝐱subscript~Ω0𝐄andsubscript𝑄𝑡subscript~Ω0𝑄\dot{\mathbf{x}}=-a\mathbf{x}+\frac{b|\Gamma|}{3}\mathbf{E}+\mathbf{g}_{ext},~% {}~{}\dot{\mathbf{E}}=(\alpha\mathbf{x}+\beta\dot{\mathbf{x}})-\tilde{\Omega}_% {0}\mathbf{E},~{}~{}\mbox{and}~{}~{}Q_{t}=-\tilde{\Omega}_{0}Qover˙ start_ARG bold_x end_ARG = - italic_a bold_x + divide start_ARG italic_b | roman_Γ | end_ARG start_ARG 3 end_ARG bold_E + bold_g start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT , over˙ start_ARG bold_E end_ARG = ( italic_α bold_x + italic_β over˙ start_ARG bold_x end_ARG ) - over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT bold_E , and italic_Q start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = - over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_Q (16)

So, for a sphere, the linearized Eqs. (12,13) can be decomposed into (i) a 6-dimensional set of ODEs (for 𝐱𝐱\mathbf{x}bold_x and 𝐄𝐄\mathbf{E}bold_E) in which each spatial direction can be treated independently and identically (modulo the direction of the external force), and which couple positional dynamics 𝐱𝐱\mathbf{x}bold_x to the evolving l=1𝑙1l=1italic_l = 1 components 𝐄𝐄\mathbf{E}bold_E of surface probability P1subscript𝑃1P_{1}italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT; (ii), an infinite number of surface harmonics evolved in Q𝑄Qitalic_Q, all of which decay to zero at rate Ω~0subscript~Ω0\tilde{\Omega}_{0}over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. We find a similar but somewhat more complex structure for the generalized spheroid case. This generalizes and clarifies the analysis of [29] who assumed an axisymmetric P0subscript𝑃0P_{0}italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and centrosome motion along the z𝑧zitalic_z-axis, and did not analyze the dynamics of the remainder modes in Q𝑄Qitalic_Q.

The linearized dynamics of the system can now be understood through two scalar ODEs (say, by projecting along the direction of the external force). From these we can obtain a second-order ODE for the linear dynamics:

η~x¨+(η~Ω~0+P03W(2κWVg))x˙+P03W(2Ω~0κ)x=Ω~0f~ext+f~˙ext,~𝜂¨𝑥~𝜂subscript~Ω0subscript𝑃03𝑊2𝜅𝑊subscript𝑉𝑔˙𝑥subscript𝑃03𝑊2subscript~Ω0𝜅𝑥subscript~Ω0subscript~𝑓𝑒𝑥𝑡subscript˙~𝑓𝑒𝑥𝑡\displaystyle\tilde{\eta}\ddot{x}+\bigg{(}\tilde{\eta}\tilde{\Omega}_{0}+\frac% {P_{0}}{3W}\left(2-\kappa\frac{W}{V_{g}}\right)\bigg{)}\dot{x}+\frac{P_{0}}{3W% }\left(2\tilde{\Omega}_{0}-\kappa\mathcal{B}\right)x=\tilde{\Omega}_{0}{\tilde% {f}}_{ext}+\dot{\tilde{f}}_{ext}\,,over~ start_ARG italic_η end_ARG over¨ start_ARG italic_x end_ARG + ( over~ start_ARG italic_η end_ARG over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + divide start_ARG italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 3 italic_W end_ARG ( 2 - italic_κ divide start_ARG italic_W end_ARG start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG ) ) over˙ start_ARG italic_x end_ARG + divide start_ARG italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 3 italic_W end_ARG ( 2 over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_κ caligraphic_B ) italic_x = over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT over~ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT + over˙ start_ARG over~ start_ARG italic_f end_ARG end_ARG start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT , (17)

where =(W/lc+rmDχ(rmD)χ(rmD))=W/lc+2+𝒪((rmD)2)𝑊subscript𝑙𝑐subscript𝑟𝑚𝐷superscript𝜒subscript𝑟𝑚𝐷𝜒subscript𝑟𝑚𝐷𝑊subscript𝑙𝑐2𝒪superscriptsubscript𝑟𝑚𝐷2{\cal B}=\left(W/l_{c}+\frac{r_{m}}{D}\frac{\chi^{\prime}(\frac{r_{m}}{D})}{% \chi(\frac{r_{m}}{D})}\right)=W/l_{c}+2+\mathcal{O}((\frac{r_{m}}{D})^{2})caligraphic_B = ( italic_W / italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT + divide start_ARG italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG italic_D end_ARG divide start_ARG italic_χ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( divide start_ARG italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG italic_D end_ARG ) end_ARG start_ARG italic_χ ( divide start_ARG italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG italic_D end_ARG ) end_ARG ) = italic_W / italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT + 2 + caligraphic_O ( ( divide start_ARG italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG italic_D end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ). To examine the linear stability of the fixed point we set f~extsubscript~𝑓𝑒𝑥𝑡\tilde{f}_{ext}over~ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT to zero, assume xeσtsimilar-to𝑥superscript𝑒𝜎𝑡x\sim e^{\sigma t}italic_x ∼ italic_e start_POSTSUPERSCRIPT italic_σ italic_t end_POSTSUPERSCRIPT, and determine the two eigenvalues σ1,2subscript𝜎12\sigma_{1,2}italic_σ start_POSTSUBSCRIPT 1 , 2 end_POSTSUBSCRIPT.

Turning to a specific case, consider the linearized dynamics for the parameter values in Table 1 and Fig. 3: for M<M1(165)𝑀annotatedsubscript𝑀1similar-toabsent165M<M_{1}\,(\sim 165)italic_M < italic_M start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( ∼ 165 ), the two eigenvalues are real and negative, resulting in stable positioning of the centrosome at the cell center (Fig. 3A&B). These two real eigenvalues collide at M=M1𝑀subscript𝑀1M=M_{1}italic_M = italic_M start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, and a complex pair of eigenvalues with negative real parts emerges, giving damped oscillations of the centrosome when perturbed from the cell center for MH>M>M1subscript𝑀𝐻𝑀subscript𝑀1M_{H}>M>M_{1}italic_M start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT > italic_M > italic_M start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. At M=MH(525)𝑀annotatedsubscript𝑀𝐻similar-toabsent525M=M_{H}\,(\sim 525)italic_M = italic_M start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT ( ∼ 525 ) the real part crosses zero, becoming positive with nonzero imaginary part, that is, this is a Hopf bifurcation. The centrosome spontaneously oscillates with frequency ωH=Mf0P0(2Ω~0κ)/3Wη0.16s1subscript𝜔𝐻𝑀subscript𝑓0subscript𝑃02subscript~Ω0𝜅3𝑊𝜂similar-to0.16superscript𝑠1\omega_{H}=\sqrt{Mf_{0}P_{0}(2\tilde{\Omega}_{0}-\kappa\mathcal{B})/3W\eta}% \sim 0.16s^{-1}italic_ω start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT = square-root start_ARG italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 2 over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_κ caligraphic_B ) / 3 italic_W italic_η end_ARG ∼ 0.16 italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT. For MH<M<M2(880)subscript𝑀𝐻𝑀annotatedsubscript𝑀2similar-toabsent880M_{H}<M<M_{2}\,(\sim 880)italic_M start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT < italic_M < italic_M start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( ∼ 880 ), the real part of the eigenvalues is positive, resulting in an oscillatory motion around the center. For M2<M<Mcsubscript𝑀2𝑀subscript𝑀𝑐M_{2}<M<M_{c}italic_M start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT < italic_M < italic_M start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, both eigenvalues are real and positive. For M>Mc𝑀subscript𝑀𝑐M>M_{c}italic_M > italic_M start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, the eigenvalues are real with different signs. Thus, the S-model exhibits almost every allowable type of linear dynamics of a second order ODE as a function of CFG number M𝑀Mitalic_M.

For illustration, Figs. 3B show examples of the nonlinear dynamics in these different regimes assuming axisymmetry of P𝑃Pitalic_P and motion strictly on the zlimit-from𝑧z-italic_z -axis. The left panel shows strict relaxation and damped oscillations. The right panel shows on-axis oscillations and strict decentering (M=2000)𝑀2000(M=2000)( italic_M = 2000 ) of the centrosome. The curves are color-coded by the intervals of M𝑀Mitalic_M in Fig. 3A.

III.2 Force versus displacement

Refer to caption
Figure 4: Stable aster centering in the sphere. (A) Force-displacement coefficient, kcsubscript𝑘𝑐k_{c}italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, as a function of CFG number M𝑀Mitalic_M for the independent CFG model (blue) and the overlap model (red) using parameter values in Table 1. At M=Mc𝑀subscript𝑀𝑐M=M_{c}italic_M = italic_M start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, kcsubscript𝑘𝑐k_{c}italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT becomes negative, and the center becomes unstable. The inset shows kcsubscript𝑘𝑐k_{c}italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for a larger range of M𝑀Mitalic_M. (B) The value of Mcsubscript𝑀𝑐M_{c}italic_M start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT as a function of dimensionless MT length (lc/Wsubscript𝑙𝑐𝑊l_{c}/Witalic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_W) and γ/κ𝛾𝜅\gamma/\kappaitalic_γ / italic_κ. In the dashed region, Mc=0subscript𝑀𝑐0M_{c}=0italic_M start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 0 and the center is unstable for any range of M𝑀Mitalic_M. The red line shows the analytical result for the stable to unstable transition. The inset shows the schematic of the simulation, where red/blue indicates bound/unbound CFGs. The star indicates parameters in Table 1.

A number of experimental studies have examined the mechanics and stability of spindle centering by exerting a controlled external force on the spindle and measuring the resulting displacement [17, 22, 33]. For example, measurements in early C. elegans embryos revealed that a micron displacement of the spindle from the cell center required forces of 10pNsimilar-toabsent10𝑝N\sim 10\;p\text{N}∼ 10 italic_p N [17]. Further, it has also been shown that in mutant embryos, where motors are partially depleted, a higher force is required to displace the spindle [17]. It was argued that this behavior arose from a combination of pulling forces from CFGs acting to decenter the centrosome, and pushing forces from MTs growing against the cell periphery acting as centering ones [17]. Nonetheless, when forces upon the centrosome were probed via laser cutting [29] no substantial contributions from pushing forces was found.

Having found stable centering with only pulling forces in the S-model, it is natural to enquire the magnitude of the centering force and its variation with model parameters. To do this we set x¨=x˙=f˙ext=0¨𝑥˙𝑥subscript˙𝑓𝑒𝑥𝑡0\ddot{x}=\dot{x}=\dot{f}_{ext}=0over¨ start_ARG italic_x end_ARG = over˙ start_ARG italic_x end_ARG = over˙ start_ARG italic_f end_ARG start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT = 0 in Eq. (17), and calculate the force-displacement coefficient, kc=fext/xsubscript𝑘𝑐subscript𝑓𝑒𝑥𝑡𝑥k_{c}=f_{ext}/xitalic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_f start_POSTSUBSCRIPT italic_e italic_x italic_t end_POSTSUBSCRIPT / italic_x:

kc=Mf0P03W(2κΩ~0).subscript𝑘𝑐𝑀subscript𝑓0subscript𝑃03𝑊2𝜅subscript~Ω0\displaystyle k_{c}=\frac{Mf_{0}P_{0}}{3W}\left(2-\frac{\kappa}{\tilde{\Omega}% _{0}}\mathcal{B}\right).italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = divide start_ARG italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 3 italic_W end_ARG ( 2 - divide start_ARG italic_κ end_ARG start_ARG over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG caligraphic_B ) . (18)

For M=200𝑀200M=200italic_M = 200 and parameter values in Table 1, kc7pN/μmsimilar-tosubscript𝑘𝑐7𝑝N𝜇𝑚k_{c}\sim 7\;p\text{N}/\mu mitalic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ∼ 7 italic_p N / italic_μ italic_m, which agrees in magnitude with measurements in C. elegans [17]. While kcsubscript𝑘𝑐k_{c}italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is often referred to as a spring coefficient the centrosome of the S-model simply does not have a simple spring-dashpot response to an applied force.

Previously, for the independent CFG S-model [29], we showed that kcsubscript𝑘𝑐k_{c}italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT increases linearly with M𝑀Mitalic_M (blue curve in Fig. 4A), in apparent disagreement with experimental observation. This follows directly from Eq. (18) by setting Ω~0subscript~Ω0{\tilde{\Omega}_{0}}over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT to a constant independent of M𝑀Mitalic_M, which is true for the independent CFG S-model, but not for the overlap model; see Eq. (15). Figure 4A (red) shows kcsubscript𝑘𝑐k_{c}italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for the overlap S-model. For low M𝑀Mitalic_M, kcsubscript𝑘𝑐k_{c}italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT increases monotonically with M𝑀Mitalic_M, consistent with the independent CFG model, but is locally convex down and reaches a maximum of kc11pN/μmsimilar-tosubscript𝑘𝑐11𝑝N𝜇𝑚k_{c}\sim 11\;p\text{N}/\mu mitalic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ∼ 11 italic_p N / italic_μ italic_m around M=495𝑀495M=495italic_M = 495. Beyond that, kcsubscript𝑘𝑐k_{c}italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT decreases monotonically with increasing M𝑀Mitalic_M (and eventually becomes negative for sufficiently large M𝑀Mitalic_M). This region of increasing kcsubscript𝑘𝑐k_{c}italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT with decreasing M𝑀Mitalic_M, due to the effect of overlap between CFGs, could be an explanation for experimental measurements in C. elegans, where a decrease in motor number increases the centering stiffness [17].

Our analysis shows that for a sufficiently large CFG number MMc𝑀subscript𝑀𝑐M\geq M_{c}italic_M ≥ italic_M start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, where one of the real eigenvalues is positive and the other is negative (Fig. 3A), kcsubscript𝑘𝑐k_{c}italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is negative (Fig. 4A), and the sphere center turns to a saddle fixed point. From Eq. (18), it is straightforward to show that kcsubscript𝑘𝑐k_{c}italic_k start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT is positive only if WlceW/lcκγ2χ(rm/W)𝑊subscript𝑙𝑐superscript𝑒𝑊subscript𝑙𝑐𝜅𝛾2𝜒subscript𝑟𝑚𝑊\frac{W}{l_{c}}e^{-W/l_{c}}\frac{\kappa}{\gamma}\leq 2\chi(r_{m}/W)divide start_ARG italic_W end_ARG start_ARG italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG italic_e start_POSTSUPERSCRIPT - italic_W / italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUPERSCRIPT divide start_ARG italic_κ end_ARG start_ARG italic_γ end_ARG ≤ 2 italic_χ ( italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT / italic_W ), where two dimensionless parameters lc/Wsubscript𝑙𝑐𝑊l_{c}/Witalic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_W and γ/κ𝛾𝜅\gamma/\kappaitalic_γ / italic_κ characterize the transition from centering to decentering. We calculated the critical value of CFG number M=Mc𝑀subscript𝑀𝑐M=M_{c}italic_M = italic_M start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, at which this transition occurs as a function of lc/Wsubscript𝑙𝑐𝑊l_{c}/Witalic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_W and γ/κ𝛾𝜅\gamma/\kappaitalic_γ / italic_κ. We found that for low lc/Wsubscript𝑙𝑐𝑊l_{c}/Witalic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_W and γ/κ𝛾𝜅\gamma/\kappaitalic_γ / italic_κ, the center is always unstable (Fig. 4B). However, for larger lc/Wsubscript𝑙𝑐𝑊l_{c}/Witalic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_W and γ/κ𝛾𝜅\gamma/\kappaitalic_γ / italic_κ, the center becomes stable for M𝑀Mitalic_M smaller than MHsubscript𝑀𝐻M_{H}italic_M start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT, and the decentering transition occurs at higher Mcsubscript𝑀𝑐M_{c}italic_M start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT (Fig. 4B).

III.3 Linear behavior in a spheroid

Refer to caption
Figure 5: Linear analysis of the S-model in spheroidal geometries. (A) Four representative cell shapes parameterized by the shape factor α𝛼\alphaitalic_α with the same surface area: prolate for α>1𝛼1\alpha>1italic_α > 1, sphere for α=1𝛼1\alpha=1italic_α = 1, and oblate for α<1𝛼1\alpha<1italic_α < 1. (B) Linear stability analysis of the centrosomal aster as a function of cell shape α𝛼\alphaitalic_α and CFG number M𝑀Mitalic_M for parameter values in Table 1. The color code denotes the nine eigenmodes and associated centrosome dynamics, where “stable” means that the real part of the eigenvalues is negative, “oscillate” means that the eigenvalue has a positive real part and a non-zero imaginary part, and “unstable” means that the real part is positive and the imaginary part is zero.

In § III.1, we show that in spherical geometry, the linear dynamics of the centrosome is determined by two eigenvalues with an independent motion in each direction. While spheres, due to their symmetries, are amenable to analytical analysis, as the reader may suspect not all biological cells are spherical (e.g., see C. elegans embryo in Fig. 1A and Supplementary Video 1). In this section, we study the linear behavior of the S-model for a class of axisymmetric spheroids centered at the origin and with the z𝑧zitalic_z-axis taken as the axis of symmetry. We show that only four eigenvalues determine the dynamics of the system- two are associated with the movement in the xy𝑥𝑦xyitalic_x italic_y-plane, and the other two along the z𝑧zitalic_z-axis.

To extend the stability analysis in § III.1 to spheroids, we substitute the ansatz 𝐱(t)=𝐯eσt𝐱𝑡𝐯superscript𝑒𝜎𝑡{\bf x}(t)={\bf v}e^{\sigma t}bold_x ( italic_t ) = bold_v italic_e start_POSTSUPERSCRIPT italic_σ italic_t end_POSTSUPERSCRIPT and P1(𝐘,𝐭)=p(𝐘)eσtsubscript𝑃1𝐘𝐭𝑝𝐘superscript𝑒𝜎𝑡P_{1}({\bf Y,t})=p({\bf Y})e^{\sigma t}italic_P start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_Y , bold_t ) = italic_p ( bold_Y ) italic_e start_POSTSUPERSCRIPT italic_σ italic_t end_POSTSUPERSCRIPT as solutions into the linearized Eqs. (12, 13) and obtain

ησ𝐯𝜂𝜎𝐯\displaystyle\eta\sigma{\bf v}italic_η italic_σ bold_v =Mf0|Γ|ΓP0(𝐘)|𝐘|(𝐈𝐘^𝐘^T)𝑑AY𝐯+Mf0|Γ|p(𝐘)𝐘^𝑑AYabsent𝑀subscript𝑓0ΓsubscriptΓsubscript𝑃0𝐘𝐘𝐈^𝐘superscript^𝐘𝑇differential-dsubscript𝐴𝑌𝐯𝑀subscript𝑓0Γ𝑝𝐘^𝐘differential-dsubscript𝐴𝑌\displaystyle=-\frac{Mf_{0}}{|\Gamma|}\int_{\Gamma}\frac{P_{0}(\mathbf{Y})}{|{% \bf Y}|}\left({\bf I}-\hat{\bf Y}\hat{\bf Y}^{T}\right)dA_{Y}\cdot{\bf v}+% \frac{Mf_{0}}{|\Gamma|}\int p({\bf Y})\hat{{\bf Y}}dA_{Y}= - divide start_ARG italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG | roman_Γ | end_ARG ∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT divide start_ARG italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) end_ARG start_ARG | bold_Y | end_ARG ( bold_I - over^ start_ARG bold_Y end_ARG over^ start_ARG bold_Y end_ARG start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ) italic_d italic_A start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ⋅ bold_v + divide start_ARG italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG | roman_Γ | end_ARG ∫ italic_p ( bold_Y ) over^ start_ARG bold_Y end_ARG italic_d italic_A start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT (19)
σp(𝐘)𝜎𝑝𝐘\displaystyle\sigma p({\bf Y})italic_σ italic_p ( bold_Y ) =κP0(𝐘)(𝐀1𝐧^+σVg𝐧^)𝐯Ω~0(𝐘)p(𝐘).absent𝜅subscript𝑃0𝐘subscript𝐀1^𝐧𝜎subscript𝑉𝑔^𝐧𝐯subscript~Ω0𝐘𝑝𝐘\displaystyle=\kappa P_{0}(\mathbf{Y})\left({\bf A}_{1}\hat{\bf n}+\frac{% \sigma}{V_{g}}\hat{\bf n}\right)\cdot{\bf v}-\tilde{\Omega}_{0}(\mathbf{Y})p({% \bf Y}).= italic_κ italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) ( bold_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT over^ start_ARG bold_n end_ARG + divide start_ARG italic_σ end_ARG start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG over^ start_ARG bold_n end_ARG ) ⋅ bold_v - over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) italic_p ( bold_Y ) . (20)

Here, we focus on the eigenvalues that correspond to oscillatory/unstable centrosome motion, i.e., 𝐯𝟎𝐯0\bf v\not=0bold_v ≠ bold_0 and σ𝜎superscript\sigma\not\in\mathbb{R}^{-}italic_σ ∉ blackboard_R start_POSTSUPERSCRIPT - end_POSTSUPERSCRIPT. For a discussion of the stable spectrum see § VI.7. By rearranging Eq. (20) we get

p(𝐘)=κP0(𝐘)σ+Ω~0(𝐘)𝐧^T(𝐀1+σVg)𝐯,𝑝𝐘𝜅subscript𝑃0𝐘𝜎subscript~Ω0𝐘superscript^𝐧𝑇subscript𝐀1𝜎subscript𝑉𝑔𝐯\displaystyle p({\bf Y})=\frac{\kappa P_{0}({\bf Y})}{\sigma+\tilde{\Omega}_{0% }({\bf Y})}\hat{\bf n}^{T}\left({\bf A}_{1}+\frac{\sigma}{V_{g}}\right)\cdot{% \bf v},italic_p ( bold_Y ) = divide start_ARG italic_κ italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) end_ARG start_ARG italic_σ + over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) end_ARG over^ start_ARG bold_n end_ARG start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ( bold_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + divide start_ARG italic_σ end_ARG start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG ) ⋅ bold_v , (21)

which we substitute to Eq. (19) and obtain the nonlinear eigenvalue equation

ησ𝐯=Mf0|Γ|[ΓP0|𝐘|(𝐈𝐘^𝐘^T)𝑑A𝐘+ΓκP0Vg𝐘^𝐧^T𝐀1+σ𝐘^𝐧^TVg(σ+Ω~0)𝑑A𝐘]𝐯.𝜂𝜎𝐯𝑀subscript𝑓0Γdelimited-[]subscriptΓsubscript𝑃0𝐘𝐈^𝐘superscript^𝐘𝑇differential-dsubscript𝐴𝐘subscriptΓ𝜅subscript𝑃0subscript𝑉𝑔^𝐘superscript^𝐧𝑇subscript𝐀1𝜎^𝐘superscript^𝐧𝑇subscript𝑉𝑔𝜎subscript~Ω0differential-dsubscript𝐴𝐘𝐯\displaystyle\eta\sigma{\bf v}=\frac{Mf_{0}}{|\Gamma|}\left[-\int_{\Gamma}% \frac{P_{0}}{|{\bf Y}|}\left({\bf I}-\hat{\bf Y}\hat{\bf Y}^{T}\right)dA_{\bf Y% }+\int_{\Gamma}\kappa P_{0}\frac{V_{g}\hat{\bf Y}\hat{\bf n}^{T}{\bf A}_{1}+% \sigma\hat{\bf Y}\hat{\bf n}^{T}}{V_{g}(\sigma+\tilde{\Omega}_{0})}dA_{\bf Y}% \right]\,{\bf v}.italic_η italic_σ bold_v = divide start_ARG italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG | roman_Γ | end_ARG [ - ∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT divide start_ARG italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG | bold_Y | end_ARG ( bold_I - over^ start_ARG bold_Y end_ARG over^ start_ARG bold_Y end_ARG start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ) italic_d italic_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT + ∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT italic_κ italic_P start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT divide start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT over^ start_ARG bold_Y end_ARG over^ start_ARG bold_n end_ARG start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_σ over^ start_ARG bold_Y end_ARG over^ start_ARG bold_n end_ARG start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT end_ARG start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ( italic_σ + over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) end_ARG italic_d italic_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT ] bold_v . (22)

Using the axial symmetry of the geometry, we recast Eq. (22) in cylindrical coordinates, with θ=atan2(y,x)𝜃atan2𝑦𝑥\theta=\text{atan}2(y,x)italic_θ = atan 2 ( italic_y , italic_x ) and z𝑧zitalic_z the coordinate along the axis of symmetry, as

ησ𝐯=𝔻1𝐯+(1σ+Ω~0(z)𝔻2(z)+σσ+Ω~0(z)𝔻3(z))𝑑z𝐯,𝜂𝜎𝐯subscript𝔻1𝐯1𝜎subscript~Ω0𝑧subscript𝔻2𝑧𝜎𝜎subscript~Ω0𝑧subscript𝔻3𝑧differential-d𝑧𝐯\displaystyle\eta\sigma{\bf v}=-\mathbb{D}_{1}{\bf v}+\int\left(\frac{1}{% \sigma+\tilde{\Omega}_{0}(z)}\mathbb{D}_{2}(z)+\frac{\sigma}{\sigma+\tilde{% \Omega}_{0}(z)}\mathbb{D}_{3}(z)\right)dz\,{\bf v},italic_η italic_σ bold_v = - blackboard_D start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT bold_v + ∫ ( divide start_ARG 1 end_ARG start_ARG italic_σ + over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_z ) end_ARG blackboard_D start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_z ) + divide start_ARG italic_σ end_ARG start_ARG italic_σ + over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_z ) end_ARG blackboard_D start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_z ) ) italic_d italic_z bold_v , (23)

where we reinterpret Ω~0subscript~Ω0\tilde{\Omega}_{0}over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT as a function of z𝑧zitalic_z. The matrices 𝔻1,2,3subscript𝔻123\mathbb{D}_{1,2,3}blackboard_D start_POSTSUBSCRIPT 1 , 2 , 3 end_POSTSUBSCRIPT are diagonal with 𝔻j1,1=𝔻j2,2superscriptsubscript𝔻𝑗11superscriptsubscript𝔻𝑗22\mathbb{D}_{j}^{1,1}=\mathbb{D}_{j}^{2,2}blackboard_D start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 , 1 end_POSTSUPERSCRIPT = blackboard_D start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 , 2 end_POSTSUPERSCRIPT with the superscript denoting the entries of the matrix. Because of this algebraic structure, we know that x^,y^,z^^𝑥^𝑦^𝑧\hat{x},\hat{y},\hat{z}over^ start_ARG italic_x end_ARG , over^ start_ARG italic_y end_ARG , over^ start_ARG italic_z end_ARG are the eigenvectors of the problem, where the first two will share the same eigenvalues. This allows us to reduce the nonlinear eigenvalue problem to the following two nonlinear algebraic equations:

ησi𝜂subscript𝜎𝑖\displaystyle\eta\sigma_{i}italic_η italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT =𝔻1i,i+(𝔻2i,i(z)σi+Ω~0(z)+σi𝔻3i,i(z)σi+Ω~0(z))𝑑z,i=1,3.formulae-sequenceabsentsubscriptsuperscript𝔻𝑖𝑖1subscriptsuperscript𝔻𝑖𝑖2𝑧subscript𝜎𝑖subscript~Ω0𝑧subscript𝜎𝑖subscriptsuperscript𝔻𝑖𝑖3𝑧subscript𝜎𝑖subscript~Ω0𝑧differential-d𝑧𝑖13\displaystyle=-\mathbb{D}^{i,i}_{1}+\int\left(\frac{\mathbb{D}^{i,i}_{2}(z)}{% \sigma_{i}+\tilde{\Omega}_{0}(z)}+\sigma_{i}\frac{\mathbb{D}^{i,i}_{3}(z)}{% \sigma_{i}+\tilde{\Omega}_{0}(z)}\right)dz,\quad i=1,3.= - blackboard_D start_POSTSUPERSCRIPT italic_i , italic_i end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + ∫ ( divide start_ARG blackboard_D start_POSTSUPERSCRIPT italic_i , italic_i end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_z ) end_ARG start_ARG italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_z ) end_ARG + italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT divide start_ARG blackboard_D start_POSTSUPERSCRIPT italic_i , italic_i end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_z ) end_ARG start_ARG italic_σ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT + over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_z ) end_ARG ) italic_d italic_z , italic_i = 1 , 3 . (24)

We numerically solve Eqs. (24) and find that there are only two solutions per index. The first two eigenvalues are associated with four eigenmodes in the xylimit-from𝑥𝑦xy-italic_x italic_y -plane and the remaining two correspond to motion in the z𝑧zitalic_z axis. Importantly, each direction of motion is completely independent of the others, just as in the case of the sphere. In § VI.6 we provide a geometric argument on the structure of these solutions.

We further compute the eigenvalues of Eqs. (24) for a family of spheroidal shapes ΓαsubscriptΓ𝛼\Gamma_{\alpha}roman_Γ start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT, defined as (x/a)2+(y/a)2+(z/c)2=1superscript𝑥𝑎2superscript𝑦𝑎2superscript𝑧𝑐21(x/a)^{2}+(y/a)^{2}+(z/c)^{2}=1( italic_x / italic_a ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( italic_y / italic_a ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( italic_z / italic_c ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1, with α=c/a𝛼𝑐𝑎\alpha=c/aitalic_α = italic_c / italic_a as the shape factor. The parameter a𝑎aitalic_a is computed such that the total area of the cell remains invariant as we vary α𝛼\alphaitalic_α (Fig. 5A). Details of the discretization and eigenvalue computation are provided in § VI.5. We construct a phase diagram as a function of CFG number M𝑀Mitalic_M and shape factor α𝛼\alphaitalic_α by classifying each point based on the eigenvalues and the corresponding centrosome movements in each direction: stable centering, oscillatory, and unstable.

For α=1𝛼1\alpha=1italic_α = 1, which is a sphere, we find three regimes that we discussed earlier: for low M𝑀Mitalic_M, the centrosome is stable in the center (case 1). As M𝑀Mitalic_M increases, the centrosome exhibits an oscillatory behavior (case 4), and for large enough M𝑀Mitalic_M, the center becomes unstable (case 9) (Fig. 5B). For oblate spheroids with α<1𝛼1\alpha<1italic_α < 1, we observe a more diverse set of behaviors where the centrosome exhibits different dynamics in different directions. In case 2, the centrosome is stable in the xy𝑥𝑦xyitalic_x italic_y-plane but oscillates in z𝑧zitalic_z; in case 6, the centrosome is unstable in z𝑧zitalic_z direction and oscillates in the xy𝑥𝑦xyitalic_x italic_y-plane, and in case 7, the centrosome is unstable in z𝑧zitalic_z and stable in the xy𝑥𝑦xyitalic_x italic_y-plane (Fig. 5B). For prolate spheroids with α>1𝛼1\alpha>1italic_α > 1, we find case 3 where the centrosome is stable in z𝑧zitalic_z but oscillates in the xy𝑥𝑦xyitalic_x italic_y-plane, a small parameter region that the centrosome is stable in z𝑧zitalic_z and unstable in the xy𝑥𝑦xyitalic_x italic_y-plane (case 8), and a region that centrosome oscillates in z𝑧zitalic_z and is unstable in the xy𝑥𝑦xyitalic_x italic_y-plane (case 5) (Fig. 5B). The diverse dynamics we observe in spheroids highlight the effect of cell geometry on centrosome motion.

III.4 Nonlinear structures of the Hopf bifurcation

Refer to caption
Figure 6: Nonlinear dynamics of the centrosome near Hopf bifurcation. (A) Oscillation amplitude as a function of M𝑀Mitalic_M for forward (red) and backward (blue) continuations using parameter values in Table 1. The dashed line indicates the value of MHsubscript𝑀𝐻M_{H}italic_M start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT at the Hopf bifurcation. (B) Oscillation amplitude as a function of M𝑀Mitalic_M for forward (red) and backward (blue) continuations using parameter values in Table 1 with γ=400[s1]𝛾400delimited-[]superscript𝑠1\gamma=400[s^{-1}]italic_γ = 400 [ italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ]. The dashed line indicates the value of M𝑀Mitalic_M at the Hopf bifurcation. (C) A zoom-in of panel B, where the dots represent the results of the simulation. (D-F) The type of the Hopf bifurcation is shown as a function of MT nucleation rate γ𝛾\gammaitalic_γ and CFG detachment rate κ𝜅\kappaitalic_κ using parameter values in Table 1 for three values of Vg=0.5,1,1.5μm/ssubscript𝑉𝑔0.511.5𝜇𝑚𝑠V_{g}=0.5,1,1.5\mu m/sitalic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = 0.5 , 1 , 1.5 italic_μ italic_m / italic_s. In the dashed region, the centrosome is stably centered. In the pink region, the centrosome exhibits supercritical Hopf bifurcation, and in the blue region, the Hopf bifurcation is subcritical. The star in E indicates parameter values of Table 1.

Motivated by the form of the unstable eigenfunction for the spherical case, we study the nature of the Hopf bifurcation in the simplest case of an axisymmetric centrosome motion (along about the z𝑧zitalic_z-axis). To distinguish between sub- and supercritical Hopf bifurcations as a function of the bifurcation parameter M𝑀Mitalic_M, we simulate the system by gradually varying M𝑀Mitalic_M around MHsubscript𝑀𝐻M_{H}italic_M start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT (the critical CFG number at the Hopf bifurcation). For a supercritical Hopf bifurcation, the transition from stationary equilibrium to oscillation is continuous in oscillation amplitude and history independent. This transition is discontinuous for a subcritical Hopf bifurcation, and the system exhibits hysteresis.

Taking advantage of the axial symmetry, we solve Eqs. (1-3) using a time-dependent spectral code (see § VI.3). We expand the occupancy probability P𝑃Pitalic_P in Legendre polynomials and integrate the truncated system of nonlinear ODEs in time, allowing for fast and accurate simulations to sweep a large parameter space. Starting the simulation with CFG number M<MH𝑀subscript𝑀𝐻M<M_{H}italic_M < italic_M start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT, we gradually increase M𝑀Mitalic_M past the Hopf bifurcation from the left (forward continuation). Once M>MH𝑀subscript𝑀𝐻M>M_{H}italic_M > italic_M start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT, we continue this simulation by decreasing M𝑀Mitalic_M past the Hopf bifurcation from the right (backward continuation). A combination of forward and backward continuation simulations allows us to compute the hysteresis and, thus, determine the nature of the Hopf bifurcation. For example, for parameters in Table 1 (star in Fig. 6E), there is no hysteresis, the Hopf bifurcation is supercritical, and the amplitude of oscillation continuously increases from zero as we pass through the Hopf bifurcation (Fig. 6A). Changing MT nucleation rate γ𝛾\gammaitalic_γ, from 300s1300superscript𝑠1300s^{-1}300 italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT to 400s1400superscript𝑠1400s^{-1}400 italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, while keeping other parameters unchanged, we find hysteresis in the model, the bifurcation becomes subcritical, and the oscillation amplitude abruptly changes passing through the bifurcation (Fig. 6B-C).

We perform a parameter sweep as a function of MT nucleation rate γ𝛾\gammaitalic_γ and CFG detachment rate κ𝜅\kappaitalic_κ to construct a phase diagram for the type of the Hopf bifurcation in the S-model (Fig. 6D-F). We find three regimes: For low values of detachment rate κ𝜅\kappaitalic_κ and nucleation rate γ𝛾\gammaitalic_γ, the centrosome does not stably oscillate (Fig. 6D-F, stripped region). However, for sufficiently large κ𝜅\kappaitalic_κ and γ𝛾\gammaitalic_γ, the centrosome oscillates. For a fixed κ𝜅\kappaitalic_κ, with increasing γ𝛾\gammaitalic_γ, the system transients to supercritical Hopf bifurcation (Fig. 6D-F, pink), followed by subcritical Hopf bifurcation (Fig. 6D-F, blue). While the structure of the phase diagrams preserves with varying Vgsubscript𝑉𝑔V_{g}italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT, the details of these transitions change.

Comparison with experiment. Previous studies in C. elegans revealed that depleting the gpr-1/2 genes, which impairs GPR-1/2 protein functionality, results in reduced cortical pulling forces and cessation of spindle oscillation [5]. A subsequent titration study [34] demonstrated a gradual decrease in spindle oscillation amplitude with the progressive depletion of gpr-1/2, identifying a critical force-generator number threshold below which spindle oscillation stops. Within the framework of the S-model, these findings align with the behavior expected from a supercritical Hopf bifurcation governing the C. elegans spindle. Our model predicts oscillation amplitudes of approximately 5μsimilar-toabsent5𝜇\sim 5\mu∼ 5 italic_μm, similar to experimental observations (5μsimilar-toabsent5𝜇\sim 5\mu∼ 5 italic_μm for maximum oscillation amplitude), and oscillation frequencies around 0.16s1similar-toabsent0.16superscript𝑠1\sim 0.16s^{-1}∼ 0.16 italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT (similar to experiment, 0.25s1similar-toabsent0.25superscript𝑠1\sim 0.25s^{-1}∼ 0.25 italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT). Notably, the model confirms the occurrence of a supercritical Hopf bifurcation. Despite the model’s simplification—comparing the C. elegans spindle, represented as two connected centrosomes in an ellipsoidal geometry, to a single centrosome within a spherical cell—the theoretical and experimental results show remarkable concordance.

IV Equatorial orbits and centrosome competition

In this section, we explore the nonlinear behavior of the system under a variety of conditions. This includes investigating aster dynamics for large values of M𝑀Mitalic_M, analyzing the impact of MT front dynamics on centrosome motion, and examining geometries with discrete rotational symmetries. We first demonstrate the emergence of periodic attractors in systems with a freely moving centrosome, positing that these attractors represent energetically preferable states compared to on-axis oscillatory motions. Next, we present evidence that stoichiometry alone can effectively position two centrosomes at diametrically opposite ends of the cell, mirroring the arrangement of spindle poles during mitosis. Finally, we demonstrate that multiple asters inside a sphere interact with each other through a competition-induced repulsion that can position themselves into symmetric configurations, as electrons constrained to a sphere and interacting via repulsive Coulomb potentials in the classical Thomson problem.

IV.1 Nonlinear orbital dynamics beyond MHsubscript𝑀𝐻M_{H}italic_M start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT

Here, we continue with the numerical analysis of the 3D nonlinear dynamics of the S-model aster, for which we will need to take account of the dynamics of the MT-front 𝐒𝐒\mathbf{S}bold_S. We examine three distinct cell shapes: prolate spheroid (α=1.5𝛼1.5\alpha=1.5italic_α = 1.5, similar to C. elegans single cell embryo), sphere (α=1𝛼1\alpha=1italic_α = 1), and oblate spheroid (α=0.5𝛼0.5\alpha=0.5italic_α = 0.5), with all taken to have equal surface areas. All simulations are for M=600>MH𝑀600subscript𝑀𝐻M=600>M_{H}italic_M = 600 > italic_M start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT (MH=325,415,525subscript𝑀𝐻325415525M_{H}=325,415,525italic_M start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT = 325 , 415 , 525, for α=0.5𝛼0.5\alpha=0.5italic_α = 0.5, 1.51.51.51.5, and 1111, respectively) and are prepared as follows. We initially fix the centrosome position at a random location 𝐗0subscript𝐗0\mathbf{X}_{0}bold_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT inside the cell with 𝐒𝐒\mathbf{S}bold_S, the MT front position a sphere of radius zero at 𝐗0subscript𝐗0\mathbf{X}_{0}bold_X start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (i.e., initially no MTs in bulk), and with P(,t=0)0𝑃𝑡00P(\cdot,t=0)\equiv 0italic_P ( ⋅ , italic_t = 0 ) ≡ 0 (all CFGs unoccupied). Holding the centrosome in place, we evolve the occupancy probability P𝑃Pitalic_P via Eq. (2) until P𝑃Pitalic_P reaches an equilibrium Psubscript𝑃P_{\infty}italic_P start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT. In the language of Eq. (1) at long times we are exerting an external force 𝐟ext=Mf0ΓP𝝃^𝑑Asubscriptsuperscript𝐟𝑒𝑥𝑡𝑀subscript𝑓0subscriptΓsubscript𝑃^𝝃differential-d𝐴\mathbf{f}^{ext}_{\infty}=-Mf_{0}\int_{\Gamma}P_{\infty}\hat{\bm{\xi}}dAbold_f start_POSTSUPERSCRIPT italic_e italic_x italic_t end_POSTSUPERSCRIPT start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = - italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT over^ start_ARG bold_italic_ξ end_ARG italic_d italic_A. We then release the centrosome to move while updating the MT front according to the algorithm described in § VI.2. We performed 10 simulations for each cell shape, and show examples in Fig. 7.

For the prolate and spherical cases, we find that the centrosome can move into an internal equatorial orbit, which can be interpreted as a steady traveling wave solution of the S-model. Consider first the prolate case (Fig. 7A). Our simulations show that, with near independence of initial position, the centrosome is attracted to the xy𝑥𝑦xyitalic_x italic_y-plane (panel Ai) and thence to a circular orbit in that plane (initial data on the z𝑧zitalic_z-axis is attracted to the origin). Different initial data yield different directions of motion upon this attractor. The colorfield in panel Aii shows the translating P𝑃Pitalic_P field for steady motion on the orbital attractor (in the plot, moving rightwards). The dashed curves show contours of the impingement rate ΩΩ\Omegaroman_Ω whose motion leads the P𝑃Pitalic_P field. Their azimuthal asymmetry reflects the mechanism that drives this traveling wave. If more MTs are impinging on the right than on the left, as they are in Fig. 7Aii & Bii, this creates a force imbalance on the centrosome that pulls it to the right, reinforcing the effect. Note that in this case the traveling wave speed along the cell surface is 1.3μm/ssimilar-toabsent1.3𝜇𝑚𝑠\sim 1.3\;\mu m/s∼ 1.3 italic_μ italic_m / italic_s, greater than the MT growth speed Vg=1μm/ssubscript𝑉𝑔1𝜇𝑚𝑠V_{g}=1\;\mu m/sitalic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT = 1 italic_μ italic_m / italic_s which means that in the aft of this wave, growing centrosomal MTs never reach the cortex (i.e. ΩΩ\Omegaroman_Ω is zero there; patterned region in Fig. 7Aii).

For the case of the sphere (Fig. 7B) we again find attraction onto internal equatorial orbits, but since the sphere geometry make no selection of a particular equator, these attracting orbits depend upon initial conditions (Fig. 7Bi). The traveling wave fields for P𝑃Pitalic_P and ΩΩ\Omegaroman_Ω are again asymmetric, closely resembling those found in the prolate case, again reflecting the asymmetry in centrosomal MT impingement that drives the wave. We have numerically constructed the equatorial orbit solution by assuming a traveling wave structure for Eqs. (1,2) which becomes a nonlinear eigenvalue-eigenfunction problem for the angular wave-speed ω𝜔\omegaitalic_ω, centrosome orbital radius ϱitalic-ϱ\varrhoitalic_ϱ, and the P𝑃Pitalic_P field, and is solved via an iterative gradient method, see § VI.4. We find good agreement between the long-time simulations and the numerical solution for exact equatorial orbits, and Fig. 7Bii shows the P𝑃Pitalic_P and ΩΩ\Omegaroman_Ω fields, as in Fig. 7Aii.

The equatorial attractors emerge essentially with the Hopf bifurcation. Figure 7Ci shows that ϱitalic-ϱ\varrhoitalic_ϱ (black curve) has a weak dependence on M𝑀Mitalic_M, as does the orbital speed ϱωitalic-ϱ𝜔\varrho\omegaitalic_ϱ italic_ω (red curve). Returning to M=600𝑀600M=600italic_M = 600, we consider the energy {\cal E}caligraphic_E for centrosomes started initially near the fixed point at cell center (panel Cii). The red curve is typical; the energy shows an immediate oscillation as the centrosome is oscillating along a line through the origin. After several oscillations the centrosome moves outwards and relaxes into an equatorial orbit of lower, and constant, energy. For comparison, the green curve shows the dynamics for a centrosome started upon the z𝑧zitalic_z-axis. Symmetries restrict its motion to oscillations along the axis, with its energy oscillating about the energy level of the fixed point itself (dashed line). Figure 7Ciii examines the behavior of {\cal E}caligraphic_E across three allowed types of dynamics: centrosome at the center, on-axis oscillation (along the z𝑧zitalic_z-axis), and moving on an equatorial orbit. Note that there is a region of bistability of the second and third cases, and that the equatorial orbit shows the lowest energy overall.

Finally, in the oblate cell, we only observed on-axis oscillation along the z𝑧zitalic_z axis. This observation is consistent with the linear stability analysis in Fig. 4B, where the only unstable mode for an oblate spheroid is along the z𝑧zitalic_z axis, while for prolate the most unstable modes are oscillations in the xy𝑥𝑦xyitalic_x italic_y-plane.

Refer to caption
Figure 7: Three-dimensional nonlinear centrosome dynamics in spheroids for parameters in Table 1. (A-B)i: Trajectories of a centrosome starting from different initial positions (black dots) in a sphere and prolate spheroid (α=1.5𝛼1.5\alpha=1.5italic_α = 1.5) with M=600𝑀600M=600italic_M = 600. (A-B)ii: The contour plots of steady-state surface occupancy probability P𝑃Pitalic_P (in color) and impingement rate ΩΩ\Omegaroman_Ω (dashed lines), as a function of azimuthal angle ϕitalic-ϕ\phiitalic_ϕ about the equator, and signed height along the z𝑧zitalic_z-axis. The patterned region shows where Ω=0s1Ω0superscript𝑠1\Omega=0~{}s^{-1}roman_Ω = 0 italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, the black dot indicates the centrosome position, and the arrow shows the orbit direction. (C)i: For case (B) of a sphere, the steady-state values of orbital radius ϱitalic-ϱ\varrhoitalic_ϱ (black) and orbital speed ϱωitalic-ϱ𝜔\varrho\omegaitalic_ϱ italic_ω (blue), as a function of M𝑀Mitalic_M. Dots show the values from full simulations, while solid lines indicate solutions found numerically from a traveling wave Ansatz. The dashed line shows M=MH𝑀subscript𝑀𝐻M=M_{H}italic_M = italic_M start_POSTSUBSCRIPT italic_H end_POSTSUBSCRIPT. (C)ii: The energy {\cal E}caligraphic_E versus time for on-axis oscillation (green) and orbital oscillations (red). The horizontal dashed line indicates the time average of the energy for on-axis oscillation. (C)iii: The time-averaged energy as a function of M𝑀Mitalic_M (green, transient solutions; red, asymptotic steady states). Both on-axis and orbital solutions are stable for M𝑀Mitalic_M between 526526526526 (dashed line) and 535535535535 (dots).

IV.2 Centrosome competition and positioning

Refer to caption
Figure 8: Arrangement of multiple centrosomes inside a sphere for parameter values in Table 1 and M=100𝑀100M=100italic_M = 100. The surface color-field in (A-C) shows the total occupancy probability. (A) Steady state occupancy probabilities for two centrosomes in a spherical cell, as a function of the polar angle θ𝜃\thetaitalic_θ from the axis that aligns with the centrosomes. The inset shows the two centrosomes (black dots) inside of the cell. (B) Stable positioning of six centrosomes, forming an octahedron, inside the sphere. Supplementary Video 2 shows the relaxation of six asters into the vertices of an octahedron. (C) Stable positioning of 3, 4, 8, and 20 centrosomes in the sphere. (D-E) The average distance of centrosomes from the sphere center (cell radius is W=15𝑊15W=15italic_W = 15) and total energy {\cal E}caligraphic_E for simulations in (A-C). Dots in E denote the total energy at the beginning of the simulation when all the centrosomes are placed near the center, bars over N=8𝑁8N=8italic_N = 8 and 20202020 represent the energy of the platonic solids, less than 1%percent11\%1 % higher than the non-platonic equilibrium configurations.

So far, we have discussed the positioning of single asters in cells. Prior to mitosis, however, the centrosome duplicates and the two newly formed asters position themselves at the two poles of the spindle, itself typically aligned along the long axis of the cell. The position and orientation of spindle poles then determine the position and orientation of the division furrow. In Farhadifar et al. [23], we used a discrete formulation of the independent motor S-model for two asters and showed that it quantitatively explains the dynamics of spindle elongation and length determination in the embryos of C. elegans and other nematodes. In a recent study, Fujii et al. created a C. elegans strain in which more than two centrosomal asters coexist within the same cell [35]. This strain is created by the deletion of the emb-27 and klp-18 genes, which yields an enucleated embryo with functional centrosomes. As the embryo progresses through its cycles, the centrosomes duplicate, but the mitotic furrow does not fully form, thus generating an embryo with multiple asters within the same cell. The authors observed that after each round of duplication, the new asters separate from each other and spread out throughout the cell.

Inspired by these experiments, we performed simulations of multiple centrosomes (N=221𝑁221N=2-21italic_N = 2 - 21) in a sphere for parameters in Table 1 and M=100𝑀100M=100italic_M = 100. For these parameter values, a single centrosome is stably centered. In our S-model, note that centrosomes only interact through their stoichiometry-mediated competition for CFGs. That is, if a CFG is occupied by MTs of centrosome A it cannot be simultaneously occupied by MTs of centrosome B. This competition creates an effective repulsive interaction between asters: If two asters are nearby then the CFGs within their common reach will divide their pulling forces between the two asters, with that division determined by the relative impingement rates of each aster. This creates a force imbalance towards the directions of noncompetitive binding, that is, the centrosomes will move away from each other. A linear calculation for the stability of two centrosomes at the sphere center confirms this is an unstable configuration for any value of M𝑀Mitalic_M, and two centrosomes always move away from the center in the opposite direction.

This is ably illustrated for the case of two centrosomes (N=2𝑁2N=2italic_N = 2). To start, we hold the centrosomes fixed near each other close to the cell center, find the corresponding steady occupancy probability using Eq. (8), and then release the asters. As Fig. 8A shows, the two centrosomes separate from each other and move to opposing sides of the sphere, with their pointwise CFG occupancy probabilities maximal near their corresponding pole and decreasing to small values further away. Each centrosome does not go all the way to the cell periphery because when it is right by its pole, the pulling force toward its pole reduces to zero and the centrosome is pulled to move toward the opposite pole. If repeated within an ellipsoidal cell, the two asters will migrate towards opposite poles of the long axis, as demonstrated in [23]. and seen in Fig. 1B.

Keeping the number of CFGs the same, we repeat this simulation with N=6𝑁6N=6italic_N = 6 asters and find that the centrosomes spread outwards and move to the vertices of a octahedron, a platonic solid, with those vertices sitting on a sphere of radius ρ6=11.63<W(=15\rho_{6}=11.63<W(=15italic_ρ start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT = 11.63 < italic_W ( = 15 μ𝜇\muitalic_μm); see Fig. 8B and Supplementary Video 2. The steady CFG occupancy probability for each centrosome is highest at the nearest cell surface point. This ultimate arrangement of asters is allowed by the spherical geometry and uniformity of the CFG distribution, as are many other possible final states (e.g. all asters sitting on the equator of an internal sphere whose radius is to be determined). That said, we find that generic initial data of centrosome positions (i.e. having no respected symmetry) relax to a platonic octahedron on a sphere of radius ρ6subscript𝜌6\rho_{6}italic_ρ start_POSTSUBSCRIPT 6 end_POSTSUBSCRIPT.

Interestingly, we find that the N=4𝑁4N=4italic_N = 4, 6666, and 12121212 cases also appear to be “globally attracted” to their corresponding platonic solid; see Figs. 8C. Not so for the N=8𝑁8N=8italic_N = 8 and 20202020 cases where the platonic solids appears to have much more limited basins of attraction, and for which we also observe other apparent steady-states. For example, for N=8𝑁8N=8italic_N = 8, if we start the simulation with asters on the vertices of a centered cube on a small sphere, it will remain a cube and expands radially until it reaches the sphere of ρ8subscript𝜌8\rho_{8}italic_ρ start_POSTSUBSCRIPT 8 end_POSTSUBSCRIPT. More generally, we instead find relaxation to the square antiprism, a twisted cube, where two opposing square faces are rotated 45454545 degrees relative to each other (see Supplementary Video 3, where we initiate dynamics from a very slightly twisted cube). For N=20𝑁20N=20italic_N = 20, with general initial conditions, we found a non-platonic configuration that resembles the 20 vertex elongated square cupola. The higher the number of centrosomes, the closer they position to the cell periphery (Fig. 8D). For other values of N𝑁Nitalic_N, the asters are almost evenly positioned upon the surface of an internal sphere, with a variance of 103superscript10310^{-3}10 start_POSTSUPERSCRIPT - 3 end_POSTSUPERSCRIPT in centrosome distance to the center. The mean centrosome distance to the center increases with N𝑁Nitalic_N (Fig. 8D), which corresponds to the plateau of total energy {\cal E}caligraphic_E at equilibrium (vertical bars in Fig. 8E), 20%percent2020\%20 % (40%percent4040\%40 %) lower than the initial energy for N=4𝑁4N=4italic_N = 4 (N=20𝑁20N=20italic_N = 20) centrosomes placed at the sphere center (diamonds in Fig. 8E).

We also investigated numerically the local stability of all the platonic solid cases (N{4,6,8,12,20}𝑁4681220N\in\{4,6,8,12,20\}italic_N ∈ { 4 , 6 , 8 , 12 , 20 }) by randomly perturbing the centrosome positions at the vertices of equilibrium platonic solids on a sphere of radius ρNsubscript𝜌𝑁\rho_{N}italic_ρ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT. We first determine ρNsubscript𝜌𝑁\rho_{N}italic_ρ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT by simulating the relaxation of centrosomes, placed on the platonic solid verticies of radius ρN0<Wsubscriptsuperscript𝜌0𝑁𝑊\rho^{0}_{N}<Witalic_ρ start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT < italic_W at the beginning of the simulations, until they reach an equilibrium distance ρNsubscript𝜌𝑁\rho_{N}italic_ρ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT to the center. Throughout these simulations the centrosome configurations retain their platonic symmetries, and relax to an equilibrium distance ρNsubscript𝜌𝑁\rho_{N}italic_ρ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT with a variance less than 106superscript10610^{-6}10 start_POSTSUPERSCRIPT - 6 end_POSTSUPERSCRIPT. We then perturb the centrosome positions from the equilibrium platonic solid configurations. For N=4𝑁4N=4italic_N = 4, 6666 and 12121212, we observe the randomly perturbed configurations relax back to the platonic solid of radii ρNsubscript𝜌𝑁\rho_{N}italic_ρ start_POSTSUBSCRIPT italic_N end_POSTSUBSCRIPT, suggesting that these are locally stable solutions. For N=8𝑁8N=8italic_N = 8 and N=20𝑁20N=20italic_N = 20, however, the perturbed configurations evolve to non-platonic shapes, suggesting that the cube (N=8𝑁8N=8italic_N = 8) and dodecahedron (N=20𝑁20N=20italic_N = 20) configurations are unstable solutions. Interestingly we also find that, for both N=8𝑁8N=8italic_N = 8 and N=20𝑁20N=20italic_N = 20, the platonic configuration has a total energy {\cal E}caligraphic_E (horizontal bars in Fig. 8E) that is slightly (less than 1%percent11\%1 %) higher than the non-platonic equilibrium configuration.

The ordering of centrosomes inside a sphere is indicative of an effective repulsion between centrosomes in the S-model. We can understand this heuristically in a simplified version of the S-model. As in Farhadifar et al. [23], assume that the binding probability is determined quasi-statically (i.e. Pti=0subscriptsuperscript𝑃𝑖𝑡0P^{i}_{t}=0italic_P start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT = 0) and that (P)=1P𝑃1𝑃{\cal I}(P)=1-Pcaligraphic_I ( italic_P ) = 1 - italic_P (i.e., the independent motor model). In this case, the occupancy probabilities can be found exactly in terms of the impingement rates as Pi(𝐘)=Ωi(𝐘)/(κ+Ω¯(𝐘))superscript𝑃𝑖𝐘superscriptΩ𝑖𝐘𝜅¯Ω𝐘P^{i}(\mathbf{Y})=\Omega^{i}(\mathbf{Y})/(\kappa+\bar{\Omega}(\mathbf{Y}))italic_P start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( bold_Y ) = roman_Ω start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( bold_Y ) / ( italic_κ + over¯ start_ARG roman_Ω end_ARG ( bold_Y ) ) where Ω¯=Σi=1NΩi¯ΩsuperscriptsubscriptΣ𝑖1𝑁superscriptΩ𝑖\bar{\Omega}=\Sigma_{i=1}^{N}\Omega^{i}over¯ start_ARG roman_Ω end_ARG = roman_Σ start_POSTSUBSCRIPT italic_i = 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT roman_Ω start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT, that is, the total impingement rate from all centrosomes at a surface point (𝐘)𝐘(\mathbf{Y})( bold_Y ). In this approximation we then have

η𝐗˙i=Mf0ΓΩi(𝐘)κ+Ω¯(𝐘)𝝃^i(𝐘)ρ(𝐘)𝑑A𝐘.𝜂superscript˙𝐗𝑖𝑀subscript𝑓0subscriptΓsuperscriptΩ𝑖𝐘𝜅¯Ω𝐘superscript^𝝃𝑖𝐘𝜌𝐘differential-dsubscript𝐴𝐘\eta\dot{\mathbf{X}}^{i}=Mf_{0}\int_{\Gamma}\frac{\Omega^{i}(\mathbf{Y})}{% \kappa+\bar{\Omega}(\mathbf{Y})}\hat{\bm{\xi}}^{i}(\mathbf{Y})\rho({\bf Y})dA_% {\mathbf{Y}}~{}.italic_η over˙ start_ARG bold_X end_ARG start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT = italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT divide start_ARG roman_Ω start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( bold_Y ) end_ARG start_ARG italic_κ + over¯ start_ARG roman_Ω end_ARG ( bold_Y ) end_ARG over^ start_ARG bold_italic_ξ end_ARG start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( bold_Y ) italic_ρ ( bold_Y ) italic_d italic_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT . (25)

Thus, the direction of centrosome motion is set by a weighted integral average of pulling force vectors f0𝝃^i(𝐘)subscript𝑓0superscript^𝝃𝑖𝐘f_{0}\hat{\bm{\xi}}^{i}(\mathbf{Y})italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT over^ start_ARG bold_italic_ξ end_ARG start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ( bold_Y ), with regions of high population impingement rate divisively reducing that region’s contribution to the total pulling force on the centrosome. Consequently, centrosomes will be pulled preferentially by their respective regions of relatively unshared CFGs. This is, in effect, a competition-induced repulsion between asters and is consistent with the linear stability calculation that shows two centrosomes at the sphere center always move away from each other.

The results found here very closely track those of the famous Thomson problem for energy-minimizing configurations of electrons constrained to a sphere and repelling each other through Coulombic forces [30]. The minimizing equilibrium configurations of up to 470470470470 electrons are available in the literature [36]. For 4444, 6666 and 12121212 electrons on a sphere, the corresponding platonic solids’ vertices (tetrahedron, octahedron, and icosahedron, respectively) are the equilibrium positions that minimize the total Coulombic potential. Eight electrons settle to the square antiprism, similar to the antiprism of eight asters in Fig. 8C but with a different aspect ratio. Twenty electrons equilibrate to the twenty vertex non-regular polyhedron with D3hsubscript𝐷3D_{3h}italic_D start_POSTSUBSCRIPT 3 italic_h end_POSTSUBSCRIPT symmetry. For both N=8𝑁8N=8italic_N = 8 and 20202020, the distribution of pairwise distances between repelling electrons is different from that between asters in Fig. 8C. For N=4𝑁4N=4italic_N = 4, 6666, and 12121212, the equilibrium relative orientations of electrons in the Thomson problem are those found here, though with the aster choosing their own sphere upon which to order. Thus, the spatial ordering of centrosomal asters inside a spherical cell, at least as modeled here, seems a new, generalized Thomson problem [37, 38], though in a nonconservative system where particles interact indirectly through competition, and move through a balance of active forces and dissipation. Finally, note that in our simulations, CFGs are uniformly distributed across the cell surface, and presumably then, due to this symmetry, the asters (nearly) evenly distribute.

V Discussion

Cortical pulling forces play a fundamental role in the positioning and orientation of the mitotic spindle, perhaps best illustrated during asymmetric division, where the spindle accurately positions off the cell center [39, 40, 41]. It has been argued that cortical pulling forces are destabilizing, and these forces alone are not sufficient to position the spindle in cells [7]. Indeed, a set of models have been developed to explain spindle positioning using mechanisms such as MT polymerization-driven pushing forces against the cell cortex, pulling forces from motor proteins carrying payloads along MTs, MT length-dependent pulling forces from cortical motors, and forces from microtubule friction with the cell wall [4, 5, 6, 7, 8, 9, 10, 11, 12, 13, 14, 15, 16, 17, 18, 19, 21, 22]. Alternatively, a simple, intuitive idea was suggested that stable centering of the aster could be accomplished by pulling forces if there are fewer CFGs than astral MTs [42]. The stoichiometric model naturally expresses this possibility through its bottom-up development and, in its earlier forms, quantitatively explains spindle positioning, elongation dynamics, and scaling with cell size among nematodes [23], and quantitatively explains the oscillatory dynamics of the spindle in C. elegans [29]. A version of the independent motor model, but with a force-dependent detachment rate, was previously used to model chromosome oscillation in human cells [8].

Here we developed an elaborated S-model that accounts for CFG binding domain overlap, which explains recent experiments, and a generalized dynamical formulation. We used this model to explore the mathematical and dynamical structure of how centrosomal asters interact with CFG populations, focusing particularly on the effects of motor density, cell shape, and centrosome number. The system is surprisingly rich in behaviors. As we describe below, this study creates the basis for the modeling of yet more complex, and fundamental, intracellular dynamics during development.

Previous in vitro and in vivo studies have demonstrated that, in some cases, aster positioning is governed by a dynamic equilibrium between the pulling forces exerted by motor proteins and the pushing forces arising from microtubule polymerization against the cell cortex [43, 44, 14, 45]. As microtubules elongate and encounter the cell periphery, they push on the centrosomal aster, potentially leading to microtubule buckling [46]. Computational tools such as Cytosim [47] and SkellySim [48, 49] have been employed to simulate these phenomena, accommodating extensive microtubule deformations and in the case of SkellySim, incorporating hydrodynamic interactions due to microtubule motion. Various coarse-grained models have also been constructed to account for microtubule pushing forces, assuming microtubules grow briefly against the cell surface and push, but disassemble before buckling  [12, 7, 15, 50]. In line with such assumptions, pushing forces can be easily accommodated within the S-model, with the local pushing force being proportional to the microtubule impingement rate ΩΩ\Omegaroman_Ω, which is part of the S-model. Out of curiosity and motivated by interesting in vitro experiments [14], we studied the positioning, via the S-model, of an aster in a flat cuboidal geometry with motors populating only the sides. While it is clear that pushing must play a role there – experimentally the aster centers in the lack of motors – we were able to reproduce substantial parts of the results solely by pulling forces (see Supplementary Fig.12).

The S-model is applicable to other biological phenomena involving aster positioning. For example, the life of C. elegans, and many other metazoans, begins with two paternally delivered centrosomal asters, which will eventually fuse the male and female pronuclei, positioned at opposite ends of the embryo. Thus begins an orchestrated movement of asters and pronuclei starting with the centrosomes’ migration toward opposing poles of the male pronucleus. Following this, the female pronucleus advances toward the male, leading to their fusion, and the resultant complex of asters and pronuclei collectively relocates to the cell center and reorients to align with the cell’s axis. Similarly, in later stages, newly duplicated centrosomes and their asters migrate to opposite sides of the nucleus and orient themselves in the direction of the future spindle. Studies have shown that SUN-KASH protein complexes sitting within the nuclear envelope can bind motor proteins such as dynein and through this will exert a pulling force on centrosomal microtubules and the aster itself. Unlike the cell cortex, which is more fixed, the nucleus is mobile and can reposition and deform due to these pulling forces. It is straightforward to expand the S-model to account for centrosome/nucleus positioning and we are doing that now. There are new issues to confront such as overall force balance, and the “shadows” cast by the nucleus that capture microtubules that might have otherwise impinged upon the cortex. Other extensions to the S-model including elastic responses such as from protein tether stretching (which can induce torques), a force/growth-velocity relationship, and hydrodynamic interactions [51].

The correspondence of multiple centrosome positioning, which is a damped and driven competition for motors, and the conservative and classical Thomson problem is amusing and surprising. We note that there have been a number of examples of damped and driven active matter systems behaving essentially as conservative systems; see, e.g. [52, 53]. While a nonuniform distribution of CFGs is important to spindle placement, we did not study it here. Such nonuniformity has been posited to explain clustering of centrosomes [54], and it would be interesting to understand what the the S-model predicts for aster arrangements in cells with non-uniform CFG distributions.

Finally, in this paper, we also examined the role of cell shape. Understanding how cell geometry regulates spindle positioning and orientation is fundamental for understanding embryonic development. For example, in C. elegans, embryogenesis progresses from a single cell encapsulated in a relatively rigid shell to about a thousand cells by successive cell divisions, leading to changes in cell volume and shape. During each division, the spindle forms and precisely orients along the long axis of the cell, often referred to as Hertwig’s rule in developmental biology [55]. The biophysical mechanism behind Hertwig’s rule has been actively pursued [56, 57, 58], and we are now testing predictions of the S-model for spindle positioning and its 3D orientation during C. elegans development, where we observed quantitative agreement with experiments.

Acknowledgments

The authors acknowledge fruitful discussions with Daniel Needleman and Jane Wang. YNY acknowledges support by the National Science Foundation (NSF) under award DMS-1951600, and of the Flatiron Institute, part of Simons Foundation. MJS acknowledges support by NSF under award DMR-2004469. Numerical computations for this work were performed at facilities supported by the Scientific Computing Core at the Flatiron Institute. The microscopy image in Fig. 1, and associated Supplementary Video 1, were obtained at the CCBScope Observatory which is part of the CCBXX{}_{\mbox{X}}start_FLOATSUBSCRIPT X end_FLOATSUBSCRIPT program.

Author Contributions

All authors conceptualized the work; YNY, VGH and MJS conducted the analysis; YNY and VGH conducted numerical simulations; VGH designed the patch-based quadrature for spheroids; VGH and HZH derived the overlap motor model; VGH, RF and YNY made the graphics; MJS supervised the research; all authors contributed to the writing.

References

  • Decker et al. [2011] M. Decker, S. Jaensch, A. Pozniakovsky, A. Zinke, K. F. O’Connell, W. Zachariae, E. Myers, and A. A. Hyman, Limiting amounts of centrosome material set centrosome size in C. elegans embryos, Current Biology 21, 1259 (2011).
  • Bornens [2012] M. Bornens, The centrosome in cells and organisms, Science 335, 422 (2012).
  • Janson and Dogterom [2004] M. E. Janson and M. Dogterom, A bending mode analysis for growing microtubules: evidence for a velocity-dependent rigidity, Biophysical Journal 87, 2723 (2004).
  • Tran et al. [2001] P. Tran, L. Marsh, V. Doye, S. Inoue, and F. Chang, A mechanism for nuclear positioning in fission yeast based on microtubule pushing, The Journal of cell biology 153, 397 (2001).
  • Grill et al. [2003] S. W. Grill, J. Howard, E. Schaffer, E. H. Stelzer, and A. A. Hyman, The distribution of active force generators controls mitotic spindle position, Science 301, 518 (2003).
  • Grill et al. [2005] S. W. Grill, K. Kruse, and F. Jülicher, Theory of mitotic spindle oscillations, Physical Review Letters 94, 108104 (2005).
  • Howard [2006] J. Howard, Elastic and damping forces generated by confined arrays of dynamic microtubules, Physical biology 3, 54 (2006).
  • Campas and Sens [2006] O. Campas and P. Sens, Centrosome oscillations in mitosis, Phys. Rev. Lett. 97, 128102 (2006).
  • Théry et al. [2007] M. Théry, A. Jiménez-Dalmaroni, V. Racine, M. Bornens, and F. Jülicher, Experimental and theoretical study of mitotic spindle orientation, Nature 447, 493 (2007).
  • Kozlowski et al. [2007] C. Kozlowski, M. Srayko, and F. Nedelec, Cortical microtubule contacts position the spindle in C𝐶\it Citalic_C. elegans embryos, Cell 129, 499 (2007).
  • Hara and Kimura [2009] Y. Hara and A. Kimura, Cell-size-dependent spindle elongation in the Caenorhabditis elegans early embryo, Current Biology 19, 1549 (2009).
  • Zhu et al. [2010] J. Zhu, A. Burakov, V. Rodionov, and A. Mogilner, Finding the cell center by a balance of dynein and myosin pulling and microtubule pushing: a computational study, Molecular biology of the cell 21, 4418 (2010).
  • Shinar et al. [2011] T. Shinar, M. Mana, F. Piano, and M. J. Shelley, A model of cytoplasmically driven microtubule-based motion in the single-celled caenorhabditis elegans embryo, Proceedings of the National Academy of Sciences 108, 10508 (2011).
  • Laan et al. [2012] L. Laan, N. Pavin, J. Husson, G. Romet-Lemonne, M. Van Duijn, M. P. López, R. D. Vale, F. Jülicher, S. L. Reck-Peterson, and M. Dogterom, Cortical dynein controls microtubule dynamics to generate pulling forces that position microtubule asters, Cell 148, 502 (2012).
  • Pavin et al. [2012] N. Pavin, L. Laan, R. Ma, M. Dogterom, and F. Jülicher, Positioning of microtubule organizing centers by cortical pushing and pulling forces, New Journal of Physics 14, 105025 (2012).
  • Ma et al. [2014a] R. Ma, L. Laan, M. Dogterom, N. Pavin, and F. Jülicher, General theory for the mechanics of confined microtubule asters, New Journal of Physics 16, 013018 (2014a).
  • Garzon-Coral et al. [2016] C. Garzon-Coral, H. A. Fantana, and J. Howard, A force-generating machinery maintains the spindle at the cell center during mitosis, Science 352, 1124 (2016).
  • Pecreaux et al. [2016] J. Pecreaux, S. Redemann, Z. Alayan, B. Mercat, S. Pastezeur, C. Garzon-Coral, A. A. Hyman, and J. Howard, The mitotic spindle in the one-cell C. elegans embryo is positioned with high precision and stability, Biophysical Journal 111, 1773 (2016).
  • Letort et al. [2016] G. Letort, F. Nedelec, L. Blanchoin, and M. Théry, Centrosome centering and decentering by microtubule network rearrangement, Molecular biology of the cell 27, 2833 (2016).
  • Coffman et al. [2016] V. C. Coffman, B. A. McDermott, B. Shtylla, and A. T. Dawes, Stronger net posterior cortical forces and asymmetric microtubule arrays produce simultaneous centration and rotation of the pronuclear complex in the early Caenorhabditis elegans embryo, Molecular Biology of the Cell 27, 3550 (2016).
  • Howard and Garzon-Coral [2017] J. Howard and C. Garzon-Coral, Physical limits on the precision of mitotic spindle positioning by microtubule pushing forces: mechanics of mitotic spindle positioning, BioEssays 39, 1700122 (2017).
  • Tanimoto et al. [2018] H. Tanimoto, J. Sallé, L. Dodin, and N. Minc, Physical forces determining the persistency and centering precision of microtubule asters, Nature physics 14, 848 (2018).
  • Farhadifar et al. [2020] R. Farhadifar, C.-H. Yu, G. Fabig, H.-Y. Wu, D. B. Stein, M. Rockman, T. Müller-Reichert, M. J. Shelley, and D. J. Needleman, Stoichiometric interactions explain spindle dynamics and scaling across 100 million years of nematode evolution, Elife 9, e55877 (2020).
  • Carminati and Stearns [1997] J. L. Carminati and T. Stearns, Microtubules orient the mitotic spindle in yeast through dynein-dependent interactions with the cell cortex, The Journal of cell biology 138, 629 (1997).
  • Kotak et al. [2012] S. Kotak, C. Busso, and P. Gönczy, Cortical dynein is critical for proper spindle positioning in human cells, Journal of Cell Biology 199, 97 (2012).
  • Gusnowski and Srayko [2011] E. M. Gusnowski and M. Srayko, Visualization of dynein-dependent microtubule gliding at the cell cortex: implications for spindle positioning, Journal of Cell Biology 194, 377 (2011).
  • Srayko et al. [2005] M. Srayko, A. Kaya, J. Stamford, and A. A. Hyman, Identification and characterization of factors required for microtubule growth and nucleation in the early C. elegans embryo, Developmental Cell 9, 223 (2005).
  • Redemann et al. [2017] S. Redemann, J. Baumgart, N. Lindow, M. Shelley, E. Nazockdast, A. Kratz, S. Prohaska, J. Brugués, S. Fürthauer, and T. Müller-Reichert, C. elegans chromosomes connect to centrosomes by anchoring into the spindle network, Nature communications 8, 15288 (2017).
  • Wu et al. [2023] H.-Y. Wu, G. Kabacaoğlu, E. Nazockdast, H.-C. Chang, M. J. Shelley, and D. J. Needleman, Laser ablation and fluid flows reveal the mechanism behind spindle and centrosome positioning, Nature Physics https://doi.org/10.1038/s41567-023-02223-z (2023).
  • Tomson [1904] J. J. Tomson, On the structure of the atom: an investigation of the stability and periods of osciletion of a number of corpuscles arranged at equal intervals around the circumference of a circle; with application of the results to the theory atomic structure, Philos. Mag. Series 6 7, 237 (1904).
  • Nazockdast et al. [2017a] E. Nazockdast, A. Rahimian, D. Zorin, and M. J. Shelley, A fast platform for simulating semi-flexible fiber suspensions applied to cell mechanics, Journal of Computational Physics 329, 173 (2017a).
  • Nazockdast et al. [2017b] E. Nazockdast, A. Rahimian, D. Needleman, and M. Shelley, Cytoplasmic flows as signatures for the mechanics of mitotic positioning, Molecular biology of the cell 28, 3261 (2017b).
  • Anjur-Dietrich et al. [2023] M. Anjur-Dietrich, V. G. Hererra, R. Farhadifar, H.-Y. Wu, H. Merta, S. Bahmanyar, M. Shelley, and D. Needleman, Clustering of cortical dynein regulates the mechanics of spindle orientation in human mitotic cells, bioRxiv , 2023 (2023).
  • Pecreaux et al. [2006] J. Pecreaux, J.-C. Roper, K. Kruse, F. Julicher, A. A. Hyman, S. W. Grill, and J. Howard, Spindle oscillations during asymmetric cell division require a threshold number of active cortical force generators, Current Biology 16, 2111 (2006).
  • Fujii et al. [2024] K. Fujii, T. Kondo, and A. Kimura, Enucleation of the C. elegans embryo revealed dynein-dependent spacing between microtubule asters, Life Science Alliance 7 (2024).
  • Atiyah and Sutcliffe [2003] M. Atiyah and P. Sutcliffe, Polyhedra in physics, chemistry and geometry, arXiv preprint math-ph/0303071  (2003).
  • Smale [1998] S. Smale, Mathematical problems for the next century, The Mathematical Intelligencer 20, 7 (1998).
  • Bowick et al. [2002] M. Bowick, A. Cacciuto, D. R. Nelson, and A. Travesset, Crystalline order on a sphere and the generalized thomson problem, Physical Review Letters 89, 185502 (2002).
  • Knoblich [2008] J. A. Knoblich, Mechanisms of asymmetric stem cell division, Cell 132, 583 (2008).
  • Gönczy [2008] P. Gönczy, Mechanisms of asymmetric cell division: flies and worms pave the way, Nature Reviews Molecular Cell Biology 9, 355 (2008).
  • Morin and Bellaïche [2011] X. Morin and Y. Bellaïche, Mitotic spindle orientation in asymmetric and symmetric cell divisions during animal development, Developmental Cell 21, 102 (2011).
  • Grill and Hyman [2005] S. W. Grill and A. A. Hyman, Spindle positioning by cortical pulling forces, Developmental Cell 8, 461 (2005).
  • Meaders et al. [2020] J. L. Meaders, S. N. de Matos, and D. R. Burgess, A pushing mechanism for microtubule aster positioning in a large cell type, Cell Reports 33 (2020).
  • Tanimoto et al. [2016] H. Tanimoto, A. Kimura, and N. Minc, Shape–motion relationships of centering microtubule asters, J. Cell Biol. 212, 777 (2016).
  • Sulerud et al. [2020] T. Sulerud, A. B. Sami, G. Li, A. Kloxin, J. Oakey, and J. Gatlin, Microtubule-dependent pushing forces contribute to long-distance aster movement and centration in xenopus laevis egg extracts, Molecular biology of the cell 31, 2791 (2020).
  • Dogterom and Yurke [1997] M. Dogterom and B. Yurke, Measurement of the force-velocity relation for growing microtubules, Science 278, 856 (1997).
  • Nedelec and Foethke [2007] F. Nedelec and D. Foethke, Collective langevin dynamics of flexible cytoskeletal fibers, New Journal of Physics 9, 427 (2007).
  • Nazockdast et al. [2017c] E. Nazockdast, A. Rahimian, D. Zorin, and M. Shelley, A fast platform for simulating semi-flexible fiber suspensions applied to cell mechanics, Journal of Computational Physics 329, 173 (2017c).
  • Dutta et al. [2024] S. Dutta, R. Farhadifar, W. Lu, G. Kabacaoğlu, R. Blackwell, D. B. Stein, M. Lakonishok, V. I. Gelfand, S. Y. Shvartsman, and M. J. Shelley, Self-organized intracellular twisters, Nature Physics , 1 (2024).
  • Ma et al. [2014b] R. Ma, L. Laan, M. Dogterom, N. Pavin, and F. Julicher, General theory for the mechanics of confined microtubule asters, N. J. Phys. 16, 013018 (2014b).
  • Stein and Shelley [2019] D. B. Stein and M. J. Shelley, Coarse graining the dynamics of immersed and driven fiber assemblies, Physical Review Fluids 4, 073302 (2019).
  • Oppenheimer et al. [2019] N. Oppenheimer, D. B. Stein, and M. J. Shelley, Rotating membrane inclusions crystallize through hydrodynamic and steric interactions, Physical Review Letters 123, 148101 (2019).
  • Shoham and Oppenheimer [2023] Y. Shoham and N. Oppenheimer, Hamiltonian dynamics and structural states of two-dimensional active particles, Physical Review Letters 131, 178301 (2023).
  • Mercadante et al. [2023] D. L. Mercadante, W. A. Aaron, S. D. Olson, and A. L. Manning, Cortical dynein drives centrosome clustering in cells with centrosome amplification, Molecular Biology of the Cell 34, ar63 (2023).
  • Hertwig [1884] O. Hertwig, Welchen Einfluß übt die Schwerkraft auf die Theilung der Zellen?, 2 (G. Fischer, 1884).
  • Pierre et al. [2016] A. Pierre, J. Sallé, M. Wühr, and N. Minc, Generic theoretical models to predict division patterns of cleaving embryos, Developmental Cell 39, 667 (2016).
  • Bosveld et al. [2016] F. Bosveld, O. Markova, B. Guirao, C. Martin, Z. Wang, A. Pierre, M. Balakireva, I. Gaugue, A. Ainslie, N. Christophorou, et al., Epithelial tricellular junctions act as interphase cell shape sensors to orient mitosis, Nature 530, 495 (2016).
  • Middelkoop et al. [2023] T. C. Middelkoop, J. Neipel, C. E. Cornell, R. Naumann, L. G. Pimpale, F. Jülicher, and S. W. Grill, A cytokinetic ring-driven cell rotation achieves Hertwig’s rule in early development, bioRxiv , 2023 (2023).
  • Redemann et al. [2010] S. Redemann, J. Pecreaux, N. W. Goehring, K. Khairy, E. H. K. Stelzer, A. A. Hyman, and J. Howard, Membrane invaginations reveal cortical sites that pull on mitotic spindles in one-cell C. elegans embryos, PLoS:ONE 5, 312301 (2010).
  • Vasil et al. [2016] G. M. Vasil, K. J. Burns, D. Lecoanet, S. Olver, B. P. Brown, and J. S. Oishi, Tensor calculus in polar coordinates using Jacobi polynomials, Journal of Computational Physics 325, 53 (2016).
  • Colbrook et al. [2021] M. Colbrook, A. Horning, and A. Townsend, Computing spectral measures of self-adjoint operators, SIAM review 63, 489 (2021).

Biophysical parameters of stoichiometric model

Biophysical Parameters Symbol Value References
spherical cell radius W𝑊Witalic_W 15151515 [μm]delimited-[]𝜇m[\mu\text{m}][ italic_μ m ] [29]
microtubule growth rate Vgsubscript𝑉𝑔V_{g}italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT 1[μms1]1delimited-[]𝜇msuperscript𝑠11\;[\mu\text{m}\;s^{-1}]1 [ italic_μ m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ] (0.51.5(0.5\sim 1.5( 0.5 ∼ 1.5 μms1)\mu\text{m}\;s^{-1})italic_μ m italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) [27]
microtubule catastrophe rate λ𝜆\lambdaitalic_λ 0.05[s1]0.05delimited-[]superscript𝑠10.05\;[s^{-1}]0.05 [ italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ] (0.0250.05s1)similar-to0.0250.05superscript𝑠1(0.025\sim 0.05\;s^{-1})( 0.025 ∼ 0.05 italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) [10]
microtubule nucleation rate γ𝛾\gammaitalic_γ 300[s1]300delimited-[]superscript𝑠1300\;[s^{-1}]300 [ italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ] (300500s1)similar-to300500superscript𝑠1(300\sim 500\;s^{-1})( 300 ∼ 500 italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) [23]
force-generator detachment rate κ𝜅\kappaitalic_κ 0.2[s1]0.2delimited-[]superscript𝑠10.2\;[s^{-1}]0.2 [ italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ] (0.10.5s1)similar-to0.10.5superscript𝑠1(0.1\sim 0.5\;s^{-1})( 0.1 ∼ 0.5 italic_s start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ) [59]
force-generator radius rmsubscript𝑟𝑚r_{m}italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT 1[μm]1delimited-[]𝜇m1\;[\mu\text{m}]1 [ italic_μ m ] (11.5(1\sim 1.5( 1 ∼ 1.5 μm)\mu\text{m})italic_μ m ) [23]
force-generator pulling force f0subscript𝑓0f_{0}italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT 10[pN]10delimited-[]𝑝N10\;[p\text{N}]10 [ italic_p N ] (510p(5\sim 10\;p( 5 ∼ 10 italic_pN) [10]
centrosome drag η𝜂\etaitalic_η 150[pNsμm1]150delimited-[]𝑝N𝑠𝜇superscriptm1150\;[p\text{N}\;s\;\mu\text{m}^{-1}]150 [ italic_p N italic_s italic_μ m start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ] (150200pNsμm1(150\sim 200\;p\text{N}\;s\;\mu\text{m}^{-1}( 150 ∼ 200 italic_p N italic_s italic_μ m start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT) [17]
Table 1: Characteristic microtubule length lc=Vg/λsubscript𝑙𝑐subscript𝑉𝑔𝜆l_{c}=V_{g}/\lambdaitalic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT / italic_λ, and number of microtubules at equilibrium N=γ/λ=γlc/Vgsubscript𝑁𝛾𝜆𝛾subscript𝑙𝑐subscript𝑉𝑔N_{\infty}=\gamma/\lambda=\gamma l_{c}/V_{g}italic_N start_POSTSUBSCRIPT ∞ end_POSTSUBSCRIPT = italic_γ / italic_λ = italic_γ italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT / italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT.

VI Supplemental Material

VI.1 Supplementary Videos

  • Supplementary Video 1: First mitotic spindle in C. elegans embryo.

  • Supplementary Video 2: Relaxation of six asters into the vertices of an octahedron inside a sphere.

  • Supplementary Video 3: Relaxation of eight asters into the vertices of a square antiprism inside a sphere. Initially placed at the vertices of a slightly rotated small cube, the centrosomes move radially first, and then rotate and relax to the vertices of a square antiprism.

VI.2 Algorithm for updating the MT front

In § II we provided a general formulation of three-dimensional MT front dynamics inside a cell surface ΓΓ\Gammaroman_Γ. Here we provide details of evolving the front of leading MT plus-ends in terms of a coordinate system centered on the centrosome with polar angle φ(t)[0,π]superscript𝜑𝑡0𝜋\varphi^{\prime}(t)\in[0,\pi]italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) ∈ [ 0 , italic_π ] and azimuthal angle θ(t)[0,2π)superscript𝜃𝑡02𝜋\theta^{\prime}(t)\in[0,2\pi)italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) ∈ [ 0 , 2 italic_π ). We focus on a convex cell surface ΓΓ\Gammaroman_Γ so there is a unique pair of coordinates (φ(t),θ(t))superscript𝜑𝑡superscript𝜃𝑡(\varphi^{\prime}(t),\theta^{\prime}(t))( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) ) that corresponds to coordinates of the force generators 𝐘𝐘{\bf Y}bold_Y on ΓΓ\Gammaroman_Γ. We assume that the centrosomal array does not rotate, and represent the location of the MT plus-end (MT front) as 𝐒=𝐒(φ(t),θ(t))=𝐗(t)+D𝐒(φ(t),θ(t))𝝃^(φ(t),θ(t))𝐒𝐒superscript𝜑𝑡superscript𝜃𝑡𝐗𝑡subscript𝐷𝐒superscript𝜑𝑡superscript𝜃𝑡^𝝃superscript𝜑𝑡superscript𝜃𝑡{\bf S}={\bf S}(\varphi^{\prime}(t),\theta^{\prime}(t))={\bf X}(t)+D_{\bf S}(% \varphi^{\prime}(t),\theta^{\prime}(t))\hat{\bm{\xi}}(\varphi^{\prime}(t),% \theta^{\prime}(t))bold_S = bold_S ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) ) = bold_X ( italic_t ) + italic_D start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) ) over^ start_ARG bold_italic_ξ end_ARG ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) ), where D𝐒subscript𝐷𝐒D_{\bf S}italic_D start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT is the distance from the centrosome to the MT front along the direction 𝝃^^𝝃\hat{\bm{\xi}}over^ start_ARG bold_italic_ξ end_ARG, pointing from the centrosome to CFGs at 𝐘(φ,θ)𝐘superscript𝜑superscript𝜃{\bf Y}(\varphi^{\prime},\theta^{\prime})bold_Y ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ) at a distance D=|𝐘𝐗|𝐷𝐘𝐗D=|{\bf Y}-{\bf X}|italic_D = | bold_Y - bold_X |. The relation between D𝐒subscript𝐷𝐒D_{\bf S}italic_D start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT and D𝐷Ditalic_D determines the state of the MT front; either

(i) the front reaches the cortex, i.e. D𝐒(φ(t),θ(t))D(φ(t),θ(t))subscript𝐷𝐒superscript𝜑𝑡superscript𝜃𝑡𝐷superscript𝜑𝑡superscript𝜃𝑡D_{\bf S}(\varphi^{\prime}(t),\theta^{\prime}(t))\geq D(\varphi^{\prime}(t),% \theta^{\prime}(t))italic_D start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) ) ≥ italic_D ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) ). Centrosomal MTs are impinging on force generators in this direction, which implies Ω(φ(t),θ(t))>0Ωsuperscript𝜑𝑡superscript𝜃𝑡0\Omega(\varphi^{\prime}(t),\theta^{\prime}(t))>0roman_Ω ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) ) > 0; or

(ii) the front grows toward the cortex, i.e. D𝐒(φ(t),θ(t))<D(φ(t),θ(t))subscript𝐷𝐒superscript𝜑𝑡superscript𝜃𝑡𝐷superscript𝜑𝑡superscript𝜃𝑡D_{\bf S}(\varphi^{\prime}(t),\theta^{\prime}(t))<D(\varphi^{\prime}(t),\theta% ^{\prime}(t))italic_D start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) ) < italic_D ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) ) and tD𝐒=Vgsubscript𝑡subscript𝐷𝐒subscript𝑉𝑔\partial_{t}D_{\bf S}=V_{g}∂ start_POSTSUBSCRIPT italic_t end_POSTSUBSCRIPT italic_D start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT = italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT. Centrosomal MTs are not impinging on force generators, which implies Ω(φ(t),θ(t))=0Ωsuperscript𝜑𝑡superscript𝜃𝑡0\Omega(\varphi^{\prime}(t),\theta^{\prime}(t))=0roman_Ω ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t ) ) = 0.

In the following we summarize the numerical algorithm for evolving the three-dimensional centrosome motion with a MT front. The superscripts denote the step in the time marching, starting from initial data of the occupancy probability P(t0)𝑃superscript𝑡0P(t^{0})italic_P ( italic_t start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ), centrosome position 𝐗(t0)𝐗superscript𝑡0{\bf X}(t^{0})bold_X ( italic_t start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ), and the unit vector 𝝃^(t0)^𝝃superscript𝑡0\hat{\bm{\xi}}(t^{0})over^ start_ARG bold_italic_ξ end_ARG ( italic_t start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ) (from the initial centrosome position to a surface point 𝐘𝐘{\bf Y}bold_Y) and the MT front position 𝐒(φ(t0),θ(t0))𝐒superscript𝜑superscript𝑡0superscript𝜃superscript𝑡0{\bf S}(\varphi^{\prime}(t^{0}),\theta^{\prime}(t^{0}))bold_S ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT 0 end_POSTSUPERSCRIPT ) ).

  1. 1.

    Given a time-step t𝑡\triangle t△ italic_t, and (𝐗(tn)({\bf X}(t^{n})( bold_X ( italic_t start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ), P(tn))P(t^{n}))italic_P ( italic_t start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) ), at time tn=ntsuperscript𝑡𝑛𝑛𝑡t^{n}=n\triangle titalic_t start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT = italic_n △ italic_t for n0𝑛0n\geq 0italic_n ≥ 0, compute 𝐗(tn+1)𝐗superscript𝑡𝑛1{\bf X}(t^{n+1})bold_X ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) as described in § II.1, and update the centrosome coordinates (φ(tn+1),θ(tn+1))superscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1(\varphi^{\prime}(t^{n+1}),\theta^{\prime}(t^{n+1}))( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ) and the distance D(φ(tn+1),θ(tn+1))𝐷superscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1D(\varphi^{\prime}(t^{n+1}),\theta^{\prime}(t^{n+1}))italic_D ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ).

  2. 2.

    Compute D~𝐒(φ(tn),θ(tn))=D𝐒n(φ(tn),θ(tn))+ΔtVgsubscript~𝐷𝐒superscript𝜑superscript𝑡𝑛superscript𝜃superscript𝑡𝑛superscriptsubscript𝐷𝐒𝑛superscript𝜑superscript𝑡𝑛superscript𝜃superscript𝑡𝑛Δ𝑡subscript𝑉𝑔\tilde{D}_{\bf S}(\varphi^{\prime}(t^{n}),\theta^{\prime}(t^{n}))=D_{\bf S}^{n% }(\varphi^{\prime}(t^{n}),\theta^{\prime}(t^{n}))+\Delta tV_{g}over~ start_ARG italic_D end_ARG start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) ) = italic_D start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) ) + roman_Δ italic_t italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT; then compute D~𝐒(φ(tn+1),θ(tn+1))subscript~𝐷𝐒superscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1\tilde{D}_{\bf S}(\varphi^{\prime}(t^{n+1}),\theta^{\prime}(t^{n+1}))over~ start_ARG italic_D end_ARG start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ) by interpolating D~𝐒(φ(tn),θ(tn))subscript~𝐷𝐒superscript𝜑superscript𝑡𝑛superscript𝜃superscript𝑡𝑛\tilde{D}_{\bf S}(\varphi^{\prime}(t^{n}),\theta^{\prime}(t^{n}))over~ start_ARG italic_D end_ARG start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) ) at the new coordinates (target points) (φ(tn+1),θ(tn+1))superscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1(\varphi^{\prime}(t^{n+1}),\theta^{\prime}(t^{n+1}))( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ).

  3. 3.

    Update D𝐒subscript𝐷𝐒D_{\bf S}italic_D start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT and the impingement rate R𝑅Ritalic_R according to the following scheme:

    • If D~𝐒(φ(tn+1),θ(tn+1))D(φ(tn+1),θ(tn+1))subscript~𝐷𝐒superscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1𝐷superscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1\tilde{D}_{\bf S}(\varphi^{\prime}(t^{n+1}),\theta^{\prime}(t^{n+1}))\geq D(% \varphi^{\prime}(t^{n+1}),\theta^{\prime}(t^{n+1}))over~ start_ARG italic_D end_ARG start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ) ≥ italic_D ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) )

      D𝐒(φ(tn+1),θ(tn+1))=D(φ(tn+1),θ(tn+1))subscript𝐷𝐒superscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1𝐷superscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1D_{\bf S}(\varphi^{\prime}(t^{n+1}),\theta^{\prime}(t^{n+1}))=D(\varphi^{% \prime}(t^{n+1}),\theta^{\prime}(t^{n+1}))italic_D start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ) = italic_D ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) )

      Ω(φ(tn+1),θ(tn+1))=[𝐕S𝐧^]+χ(rmD)(γVgeD/lc)Ωsuperscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1subscriptdelimited-[]subscript𝐕𝑆^𝐧𝜒subscript𝑟𝑚𝐷𝛾subscript𝑉𝑔superscript𝑒𝐷subscript𝑙𝑐\Omega(\varphi^{\prime}(t^{n+1}),\theta^{\prime}(t^{n+1}))=\left[{\bf V}_{S}% \cdot\hat{\bf n}\right]_{+}\chi\left(\frac{r_{m}}{D}\right)\left(\frac{\gamma}% {V_{g}}e^{-D/l_{c}}\right)roman_Ω ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ) = [ bold_V start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ⋅ over^ start_ARG bold_n end_ARG ] start_POSTSUBSCRIPT + end_POSTSUBSCRIPT italic_χ ( divide start_ARG italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG italic_D end_ARG ) ( divide start_ARG italic_γ end_ARG start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG italic_e start_POSTSUPERSCRIPT - italic_D / italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUPERSCRIPT )

    • Otherwise (D~𝐒(φ(tn+1),θ(tn+1))<D(φ(tn+1),θ(tn+1))subscript~𝐷𝐒superscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1𝐷superscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1\tilde{D}_{\bf S}(\varphi^{\prime}(t^{n+1}),\theta^{\prime}(t^{n+1}))<D(% \varphi^{\prime}(t^{n+1}),\theta^{\prime}(t^{n+1}))over~ start_ARG italic_D end_ARG start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ) < italic_D ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ))

      D𝐒(φ(tn+1),θ(tn+1))=D~𝐒(φ(tn+1),θ(tn+1))subscript𝐷𝐒superscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1subscript~𝐷𝐒superscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1D_{\bf S}(\varphi^{\prime}(t^{n+1}),\theta^{\prime}(t^{n+1}))=\tilde{D}_{\bf S% }(\varphi^{\prime}(t^{n+1}),\theta^{\prime}(t^{n+1}))italic_D start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ) = over~ start_ARG italic_D end_ARG start_POSTSUBSCRIPT bold_S end_POSTSUBSCRIPT ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) )

      Ω(φ(tn+1),θ(tn+1))=0Ωsuperscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛10\Omega(\varphi^{\prime}(t^{n+1}),\theta^{\prime}(t^{n+1}))=0roman_Ω ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ) = 0

We then update the MT front position to 𝐒(φ(tn+1),θ(tn+1))𝐒superscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1{\bf S}(\varphi^{\prime}(t^{n+1}),\theta^{\prime}(t^{n+1}))bold_S ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ), and use the updated impingement rate Ω(φ(tn+1),θ(tn+1))Ωsuperscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1\Omega(\varphi^{\prime}(t^{n+1}),\theta^{\prime}(t^{n+1}))roman_Ω ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ) to compute Pn+1superscript𝑃𝑛1P^{n+1}italic_P start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT as described in § II.1:

P(tn+1)P(tn)Δt𝑃superscript𝑡𝑛1𝑃superscript𝑡𝑛Δ𝑡\displaystyle\frac{P(t^{n+1})-P(t^{n})}{\Delta t}divide start_ARG italic_P ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) - italic_P ( italic_t start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) end_ARG start_ARG roman_Δ italic_t end_ARG =Ω(φ(tn+1),θ(tn+1))(P(tn))κP(tn+1),absentΩsuperscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1𝑃superscript𝑡𝑛𝜅𝑃superscript𝑡𝑛1\displaystyle=\Omega(\varphi^{\prime}(t^{n+1}),\theta^{\prime}(t^{n+1})){\cal I% }(P(t^{n}))-\kappa P(t^{n+1}),= roman_Ω ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ) caligraphic_I ( italic_P ( italic_t start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) ) - italic_κ italic_P ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ,
P(tn+1)𝑃superscript𝑡𝑛1\displaystyle P(t^{n+1})italic_P ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) =P(tn)+Δt1+κΔt(Ω(φ(tn+1),θ(tn+1))(P(tn))κP(tn)).absent𝑃superscript𝑡𝑛Δ𝑡1𝜅Δ𝑡Ωsuperscript𝜑superscript𝑡𝑛1superscript𝜃superscript𝑡𝑛1𝑃superscript𝑡𝑛𝜅𝑃superscript𝑡𝑛\displaystyle=P(t^{n})+\frac{\Delta t}{1+\kappa\Delta t}\left(\Omega(\varphi^{% \prime}(t^{n+1}),\theta^{\prime}(t^{n+1})){\cal I}(P(t^{n}))-\kappa P(t^{n})% \right).= italic_P ( italic_t start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) + divide start_ARG roman_Δ italic_t end_ARG start_ARG 1 + italic_κ roman_Δ italic_t end_ARG ( roman_Ω ( italic_φ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) , italic_θ start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_t start_POSTSUPERSCRIPT italic_n + 1 end_POSTSUPERSCRIPT ) ) caligraphic_I ( italic_P ( italic_t start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) ) - italic_κ italic_P ( italic_t start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) ) . (26)

This completes one time-step, and the iteration repeats until the end of the simulation.

VI.3 Spectral formulation for the axi-symmetric centrosome dynamics in sphere

Here we recast Eqs (1-3) with axisymmetry (z𝑧zitalic_z-axis as the axis of symmetry) in Legendre polynomials. The centrosome moves along the z𝑧zitalic_z-axis with a position z(t)𝑧𝑡z(t)italic_z ( italic_t ), and the occupancy probability is a function of the polar angle φ[0,π]𝜑0𝜋\varphi\in[0,\pi]italic_φ ∈ [ 0 , italic_π ]. We set the areal density ρ(𝐘)=1/|Γ|𝜌𝐘1Γ\rho\left({\bf Y}\right)=1/|\Gamma|italic_ρ ( bold_Y ) = 1 / | roman_Γ |. The equations of motion for the axisymmetric system are

ηdzdt𝜂𝑑𝑧𝑑𝑡\displaystyle\eta\frac{dz}{dt}italic_η divide start_ARG italic_d italic_z end_ARG start_ARG italic_d italic_t end_ARG =Mf020π𝑑φsinφWcosφzD(φ,z)P(φ,t),absent𝑀subscript𝑓02subscriptsuperscript𝜋0differential-d𝜑𝜑𝑊𝜑𝑧𝐷𝜑𝑧𝑃𝜑𝑡\displaystyle=\frac{Mf_{0}}{2}\int^{\pi}_{0}d\varphi\sin\varphi\frac{W\cos% \varphi-z}{D(\varphi,z)}P(\varphi,t),= divide start_ARG italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG ∫ start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_d italic_φ roman_sin italic_φ divide start_ARG italic_W roman_cos italic_φ - italic_z end_ARG start_ARG italic_D ( italic_φ , italic_z ) end_ARG italic_P ( italic_φ , italic_t ) , (27)
P(φ,t)t𝑃𝜑𝑡𝑡\displaystyle\frac{\partial P(\varphi,t)}{\partial t}divide start_ARG ∂ italic_P ( italic_φ , italic_t ) end_ARG start_ARG ∂ italic_t end_ARG =Ω(φ,z)[1eMM0(1P(φ,t))MM0]κP(φ,t),absentΩ𝜑𝑧delimited-[]1superscript𝑒𝑀subscript𝑀01𝑃𝜑𝑡𝑀subscript𝑀0𝜅𝑃𝜑𝑡\displaystyle=\Omega(\varphi,z)\left[\frac{1-e^{-\frac{M}{M_{0}}(1-P(\varphi,t% ))}}{\frac{M}{M_{0}}}\right]-\kappa P(\varphi,t),= roman_Ω ( italic_φ , italic_z ) [ divide start_ARG 1 - italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_M end_ARG start_ARG italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ( 1 - italic_P ( italic_φ , italic_t ) ) end_POSTSUPERSCRIPT end_ARG start_ARG divide start_ARG italic_M end_ARG start_ARG italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG end_ARG ] - italic_κ italic_P ( italic_φ , italic_t ) , (28)
Ω(φ,z)Ω𝜑𝑧\displaystyle\Omega(\varphi,z)roman_Ω ( italic_φ , italic_z ) =[z˙cosφ+Vg(Wzcosφ)D(φ,z)]+eD(φ,z)lcχ(rmD(φ,z))γVg,absentsubscriptdelimited-[]˙𝑧𝜑subscript𝑉𝑔𝑊𝑧𝜑𝐷𝜑𝑧superscript𝑒𝐷𝜑𝑧subscript𝑙𝑐𝜒subscript𝑟𝑚𝐷𝜑𝑧𝛾subscript𝑉𝑔\displaystyle=\left[\dot{z}\cos\varphi+\frac{V_{g}\left(W-z\cos\varphi\right)}% {D(\varphi,z)}\right]_{+}e^{-\frac{D(\varphi,z)}{l_{c}}}\chi\left(\frac{r_{m}}% {D(\varphi,z)}\right)\frac{\gamma}{V_{g}},= [ over˙ start_ARG italic_z end_ARG roman_cos italic_φ + divide start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ( italic_W - italic_z roman_cos italic_φ ) end_ARG start_ARG italic_D ( italic_φ , italic_z ) end_ARG ] start_POSTSUBSCRIPT + end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_D ( italic_φ , italic_z ) end_ARG start_ARG italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG end_POSTSUPERSCRIPT italic_χ ( divide start_ARG italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG italic_D ( italic_φ , italic_z ) end_ARG ) divide start_ARG italic_γ end_ARG start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG , (29)

where D(φ,z)=W22Wzcosφ+z2𝐷𝜑𝑧superscript𝑊22𝑊𝑧𝜑superscript𝑧2D(\varphi,z)=\sqrt{W^{2}-2Wz\cos\varphi+z^{2}}italic_D ( italic_φ , italic_z ) = square-root start_ARG italic_W start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 2 italic_W italic_z roman_cos italic_φ + italic_z start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG. We expand the occupancy probability P(φ,t)𝑃𝜑𝑡P(\varphi,t)italic_P ( italic_φ , italic_t ) in Legendre polynomials Pn(x)subscript𝑃𝑛𝑥P_{n}(x)italic_P start_POSTSUBSCRIPT italic_n end_POSTSUBSCRIPT ( italic_x ) as P(φ,t)=i=0ci(t)Pi(cosφ)𝑃𝜑𝑡subscriptsuperscript𝑖0subscript𝑐𝑖𝑡subscript𝑃𝑖𝜑P(\varphi,t)=\sum^{\infty}_{i=0}c_{i}(t)P_{i}(\cos\varphi)italic_P ( italic_φ , italic_t ) = ∑ start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ) italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( roman_cos italic_φ ), where Pi(x=cosφ)subscript𝑃𝑖𝑥𝜑P_{i}(x=\cos\varphi)italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x = roman_cos italic_φ ) is the i𝑖iitalic_ith order Legendre polynomial. To recast equations (27)-(28) into a system of nonlinear equations in z(t)𝑧𝑡z(t)italic_z ( italic_t ) and the coefficients ci(t)subscript𝑐𝑖𝑡c_{i}(t)italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_t ), we first expand the impingement rate in Legendre polynomials as

Ω(x,z)=(dzdtf1(x,z)+f2(x,z))γVg=i=0(dzdtf1i+f2i)Pi(x)γVg,Ω𝑥𝑧𝑑𝑧𝑑𝑡subscript𝑓1𝑥𝑧subscript𝑓2𝑥𝑧𝛾subscript𝑉𝑔subscriptsuperscript𝑖0𝑑𝑧𝑑𝑡subscript𝑓1𝑖subscript𝑓2𝑖subscript𝑃𝑖𝑥𝛾subscript𝑉𝑔\Omega(x,z)=\left(\frac{dz}{dt}f_{1}(x,z)+f_{2}(x,z)\right)\frac{\gamma}{V_{g}% }=\sum^{\infty}_{i=0}\left(\frac{dz}{dt}f_{1i}+f_{2i}\right)P_{i}(x)\frac{% \gamma}{V_{g}},roman_Ω ( italic_x , italic_z ) = ( divide start_ARG italic_d italic_z end_ARG start_ARG italic_d italic_t end_ARG italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_x , italic_z ) + italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_x , italic_z ) ) divide start_ARG italic_γ end_ARG start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG = ∑ start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT ( divide start_ARG italic_d italic_z end_ARG start_ARG italic_d italic_t end_ARG italic_f start_POSTSUBSCRIPT 1 italic_i end_POSTSUBSCRIPT + italic_f start_POSTSUBSCRIPT 2 italic_i end_POSTSUBSCRIPT ) italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x ) divide start_ARG italic_γ end_ARG start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG ,

where Pi(x)subscript𝑃𝑖𝑥P_{i}(x)italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x ) is the i𝑖iitalic_ith order Legendre polynomial, and

f1subscript𝑓1\displaystyle f_{1}italic_f start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT xeDlcχ(rmD)i=0f1iPi(x),absent𝑥superscript𝑒𝐷subscript𝑙𝑐𝜒subscript𝑟𝑚𝐷subscriptsuperscript𝑖0subscript𝑓1𝑖subscript𝑃𝑖𝑥\displaystyle\equiv xe^{-\frac{D}{l_{c}}}\chi\left(\frac{r_{m}}{D}\right)% \equiv\sum^{\infty}_{i=0}f_{1i}P_{i}(x),≡ italic_x italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_D end_ARG start_ARG italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG end_POSTSUPERSCRIPT italic_χ ( divide start_ARG italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG italic_D end_ARG ) ≡ ∑ start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT 1 italic_i end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x ) , (30)
f2subscript𝑓2\displaystyle f_{2}italic_f start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT Vg(Wzx)DeDlcχ(rmD)i=0f2iPi(x).absentsubscript𝑉𝑔𝑊𝑧𝑥𝐷superscript𝑒𝐷subscript𝑙𝑐𝜒subscript𝑟𝑚𝐷subscriptsuperscript𝑖0subscript𝑓2𝑖subscript𝑃𝑖𝑥\displaystyle\equiv\frac{V_{g}(W-zx)}{D}e^{-\frac{D}{l_{c}}}\chi\left(\frac{r_% {m}}{D}\right)\equiv\sum^{\infty}_{i=0}f_{2i}P_{i}(x).≡ divide start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT ( italic_W - italic_z italic_x ) end_ARG start_ARG italic_D end_ARG italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_D end_ARG start_ARG italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG end_POSTSUPERSCRIPT italic_χ ( divide start_ARG italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG italic_D end_ARG ) ≡ ∑ start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_i = 0 end_POSTSUBSCRIPT italic_f start_POSTSUBSCRIPT 2 italic_i end_POSTSUBSCRIPT italic_P start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( italic_x ) . (31)

Here we truncate at c1subscript𝑐1c_{1}italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT to obtain a system of nonlinear equations for z(t)𝑧𝑡z(t)italic_z ( italic_t ), c0(t)subscript𝑐0𝑡c_{0}(t)italic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_t ) and c1(t)subscript𝑐1𝑡c_{1}(t)italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_t ):

ηdzdt𝜂𝑑𝑧𝑑𝑡\displaystyle\eta\frac{dz}{dt}italic_η divide start_ARG italic_d italic_z end_ARG start_ARG italic_d italic_t end_ARG =Mf02[43c0zW+(2325(zW)2)c1],absent𝑀subscript𝑓02delimited-[]43subscript𝑐0𝑧𝑊2325superscript𝑧𝑊2subscript𝑐1\displaystyle=\frac{Mf_{0}}{2}\left[-\frac{4}{3}c_{0}\frac{z}{W}+\left(\frac{2% }{3}-\frac{2}{5}\left(\frac{z}{W}\right)^{2}\right)c_{1}\right],= divide start_ARG italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG [ - divide start_ARG 4 end_ARG start_ARG 3 end_ARG italic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT divide start_ARG italic_z end_ARG start_ARG italic_W end_ARG + ( divide start_ARG 2 end_ARG start_ARG 3 end_ARG - divide start_ARG 2 end_ARG start_ARG 5 end_ARG ( divide start_ARG italic_z end_ARG start_ARG italic_W end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ] , (32)
dc0dt𝑑subscript𝑐0𝑑𝑡\displaystyle\frac{dc_{0}}{dt}divide start_ARG italic_d italic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG =Ω0κc0(c0R0+13c1R1),absentsubscriptΩ0𝜅subscript𝑐0subscript𝑐0subscript𝑅013subscript𝑐1subscript𝑅1\displaystyle=\Omega_{0}-\kappa c_{0}-\left(c_{0}R_{0}+\frac{1}{3}c_{1}R_{1}% \right),= roman_Ω start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - italic_κ italic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT - ( italic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + divide start_ARG 1 end_ARG start_ARG 3 end_ARG italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) , (33)
dc1dt𝑑subscript𝑐1𝑑𝑡\displaystyle\frac{dc_{1}}{dt}divide start_ARG italic_d italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG =Ω1κc1(c0R1+c1R0+25c1R2),absentsubscriptΩ1𝜅subscript𝑐1subscript𝑐0subscript𝑅1subscript𝑐1subscript𝑅025subscript𝑐1subscript𝑅2\displaystyle=\Omega_{1}-\kappa c_{1}-\left(c_{0}R_{1}+c_{1}R_{0}+\frac{2}{5}c% _{1}R_{2}\right),= roman_Ω start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - italic_κ italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT - ( italic_c start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + divide start_ARG 2 end_ARG start_ARG 5 end_ARG italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_R start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) , (34)

with Ωi(dzdtf1i+f2i)γVgsubscriptΩ𝑖𝑑𝑧𝑑𝑡subscript𝑓1𝑖subscript𝑓2𝑖𝛾subscript𝑉𝑔\Omega_{i}\equiv\left(\frac{dz}{dt}f_{1i}+f_{2i}\right)\frac{\gamma}{V_{g}}roman_Ω start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ≡ ( divide start_ARG italic_d italic_z end_ARG start_ARG italic_d italic_t end_ARG italic_f start_POSTSUBSCRIPT 1 italic_i end_POSTSUBSCRIPT + italic_f start_POSTSUBSCRIPT 2 italic_i end_POSTSUBSCRIPT ) divide start_ARG italic_γ end_ARG start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG. For i=0𝑖0i=0italic_i = 0 and i2𝑖2i\geq 2italic_i ≥ 2, the linearized equation of cisubscript𝑐𝑖c_{i}italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT simply yields decaying dynamics:

dcidt𝑑subscript𝑐𝑖𝑑𝑡\displaystyle\frac{dc_{i}}{dt}divide start_ARG italic_d italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG start_ARG italic_d italic_t end_ARG =(κ+I0(P¯))ci,absent𝜅subscript𝐼0¯𝑃subscript𝑐𝑖\displaystyle=-\left(\kappa+I_{0}\left({\bar{P}}\right)\right)c_{i},= - ( italic_κ + italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( over¯ start_ARG italic_P end_ARG ) ) italic_c start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , (35)
I0(P¯)subscript𝐼0¯𝑃\displaystyle I_{0}\left({\bar{P}}\right)italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( over¯ start_ARG italic_P end_ARG ) =eW/lcχ(rmW)γ1eMM0(1P¯)MM0>0,absentsuperscript𝑒𝑊subscript𝑙𝑐𝜒subscript𝑟𝑚𝑊𝛾1superscript𝑒𝑀subscript𝑀01¯𝑃𝑀subscript𝑀00\displaystyle=e^{-W/l_{c}}\chi\left(\frac{r_{m}}{W}\right)\gamma\frac{1-e^{-% \frac{M}{M_{0}}(1-{\bar{P}})}}{\frac{M}{M_{0}}}>0,= italic_e start_POSTSUPERSCRIPT - italic_W / italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_χ ( divide start_ARG italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG italic_W end_ARG ) italic_γ divide start_ARG 1 - italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_M end_ARG start_ARG italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ( 1 - over¯ start_ARG italic_P end_ARG ) end_POSTSUPERSCRIPT end_ARG start_ARG divide start_ARG italic_M end_ARG start_ARG italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG end_ARG > 0 , (36)

where P¯¯𝑃{\bar{P}}over¯ start_ARG italic_P end_ARG, the occupancy probability of the base state, satisfies the equation P¯=I0(P¯)I0(P¯)+κ¯𝑃subscript𝐼0¯𝑃subscript𝐼0¯𝑃𝜅{\bar{P}}=\frac{I_{0}\left({\bar{P}}\right)}{I_{0}\left({\bar{P}}\right)+\kappa}over¯ start_ARG italic_P end_ARG = divide start_ARG italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( over¯ start_ARG italic_P end_ARG ) end_ARG start_ARG italic_I start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( over¯ start_ARG italic_P end_ARG ) + italic_κ end_ARG.  These results show that the linear instability of a centrosome in a sphere in § III.1 involves only c1subscript𝑐1c_{1}italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and z𝑧zitalic_z, and all the other modes are linearly stable and do not contribute to the instability.

VI.4 Centrosome orbiting around sphere center

Here we focus on a centrosome orbiting at a constant speed and a constant distance to the sphere center. In spherical coordinates, with the origin at the sphere center, the centrosome position 𝐗c=ϱ(t)(sinθc(t)cosϕc(t)𝐞^1+sinθc(t)sinϕc(t)𝐞^2+cosθc(t)𝐞^3)subscript𝐗𝑐italic-ϱ𝑡subscript𝜃𝑐𝑡subscriptitalic-ϕ𝑐𝑡subscript^𝐞1subscript𝜃𝑐𝑡subscriptitalic-ϕ𝑐𝑡subscript^𝐞2subscript𝜃𝑐𝑡subscript^𝐞3{\bf X}_{c}=\varrho(t)(\sin\theta_{c}(t)\cos\phi_{c}(t)\hat{\bf e}_{1}+\sin% \theta_{c}(t)\sin\phi_{c}(t)\hat{\bf e}_{2}+\cos\theta_{c}(t)\hat{\bf e}_{3})bold_X start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_ϱ ( italic_t ) ( roman_sin italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t ) roman_cos italic_ϕ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t ) over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + roman_sin italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t ) roman_sin italic_ϕ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t ) over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + roman_cos italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( italic_t ) over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) and the surface of the sphere of radius W𝑊Witalic_W has coordinates 𝐘=W(sinθcosϕ𝐞^1+sinθsinϕ𝐞^2+cosθ𝐞^3)𝐘𝑊𝜃italic-ϕsubscript^𝐞1𝜃italic-ϕsubscript^𝐞2𝜃subscript^𝐞3{\bf Y}=W(\sin\theta\cos\phi\hat{\bf e}_{1}+\sin\theta\sin\phi\hat{\bf e}_{2}+% \cos\theta\hat{\bf e}_{3})bold_Y = italic_W ( roman_sin italic_θ roman_cos italic_ϕ over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + roman_sin italic_θ roman_sin italic_ϕ over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + roman_cos italic_θ over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ), where the polar angles θc,θ[0,π]subscript𝜃𝑐𝜃0𝜋\theta_{c},\theta\in[0,\pi]italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT , italic_θ ∈ [ 0 , italic_π ], and the azimuthal angles ϕc,ϕ[π,π]subscriptitalic-ϕ𝑐italic-ϕ𝜋𝜋\phi_{c},\phi\in[-\pi,\pi]italic_ϕ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT , italic_ϕ ∈ [ - italic_π , italic_π ]. The governing equations for the centrosome dynamics in spherical coordinates (with areal density ρ(𝐘)=1/|Γ|𝜌𝐘1Γ\rho\left({\bf Y}\right)=1/|\Gamma|italic_ρ ( bold_Y ) = 1 / | roman_Γ |) are

ηϱ˙𝜂˙italic-ϱ\displaystyle\eta\dot{\varrho}italic_η over˙ start_ARG italic_ϱ end_ARG =Mf0|Γ|P(θ,ϕ,t)D(ϱ+Wcosθcosθc+Wsinθsinθccos(ϕϕc))W2sinθdθdϕ,absent𝑀subscript𝑓0Γ𝑃𝜃italic-ϕ𝑡𝐷italic-ϱ𝑊𝜃subscript𝜃𝑐𝑊𝜃subscript𝜃𝑐italic-ϕsubscriptitalic-ϕ𝑐superscript𝑊2𝜃𝑑𝜃𝑑italic-ϕ\displaystyle=\frac{Mf_{0}}{|\Gamma|}\int\frac{P(\theta,\phi,t)}{D}\left(-% \varrho+W\cos\theta\cos\theta_{c}+W\sin\theta\sin\theta_{c}\cos(\phi-\phi_{c})% \right)W^{2}\sin\theta d\theta d\phi,= divide start_ARG italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG | roman_Γ | end_ARG ∫ divide start_ARG italic_P ( italic_θ , italic_ϕ , italic_t ) end_ARG start_ARG italic_D end_ARG ( - italic_ϱ + italic_W roman_cos italic_θ roman_cos italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT + italic_W roman_sin italic_θ roman_sin italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT roman_cos ( italic_ϕ - italic_ϕ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) ) italic_W start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_sin italic_θ italic_d italic_θ italic_d italic_ϕ , (37)
ηϱθ˙c𝜂italic-ϱsubscript˙𝜃𝑐\displaystyle\eta\varrho\dot{\theta}_{c}italic_η italic_ϱ over˙ start_ARG italic_θ end_ARG start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT =Mf0AP(θ,ϕ,t)D(cosθcsinθcos(ϕϕc)cosθsinθc)W3sinθdθdϕ,absent𝑀subscript𝑓0𝐴𝑃𝜃italic-ϕ𝑡𝐷subscript𝜃𝑐𝜃italic-ϕsubscriptitalic-ϕ𝑐𝜃subscript𝜃𝑐superscript𝑊3𝜃𝑑𝜃𝑑italic-ϕ\displaystyle=\frac{Mf_{0}}{A}\int\frac{P(\theta,\phi,t)}{D}\left(\cos\theta_{% c}\sin\theta\cos(\phi-\phi_{c})-\cos\theta\sin\theta_{c}\right)W^{3}\sin\theta d% \theta d\phi,= divide start_ARG italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_A end_ARG ∫ divide start_ARG italic_P ( italic_θ , italic_ϕ , italic_t ) end_ARG start_ARG italic_D end_ARG ( roman_cos italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT roman_sin italic_θ roman_cos ( italic_ϕ - italic_ϕ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) - roman_cos italic_θ roman_sin italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) italic_W start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_sin italic_θ italic_d italic_θ italic_d italic_ϕ , (38)
ηϱsinθcϕ˙c𝜂italic-ϱsubscript𝜃𝑐subscript˙italic-ϕ𝑐\displaystyle\eta\varrho\sin\theta_{c}\dot{\phi}_{c}italic_η italic_ϱ roman_sin italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT over˙ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT =Mf0AP(θ,ϕ,t)Dsinθsin(ϕϕc)W3sinθdθdϕ,absent𝑀subscript𝑓0𝐴𝑃𝜃italic-ϕ𝑡𝐷𝜃italic-ϕsubscriptitalic-ϕ𝑐superscript𝑊3𝜃𝑑𝜃𝑑italic-ϕ\displaystyle=\frac{Mf_{0}}{A}\int\frac{P(\theta,\phi,t)}{D}\sin\theta\sin(% \phi-\phi_{c})W^{3}\sin\theta d\theta d\phi,= divide start_ARG italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_A end_ARG ∫ divide start_ARG italic_P ( italic_θ , italic_ϕ , italic_t ) end_ARG start_ARG italic_D end_ARG roman_sin italic_θ roman_sin ( italic_ϕ - italic_ϕ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) italic_W start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_sin italic_θ italic_d italic_θ italic_d italic_ϕ , (39)
Pt𝑃𝑡\displaystyle\frac{\partial P}{\partial t}divide start_ARG ∂ italic_P end_ARG start_ARG ∂ italic_t end_ARG =[ϱ˙(cosθcosθc+sinθsinθccos(ϕϕc))+\displaystyle=\left[\dot{\varrho}\left(\cos\theta\cos\theta_{c}+\sin\theta\sin% \theta_{c}\cos(\phi-\phi_{c})\right)+\right.= [ over˙ start_ARG italic_ϱ end_ARG ( roman_cos italic_θ roman_cos italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT + roman_sin italic_θ roman_sin italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT roman_cos ( italic_ϕ - italic_ϕ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) ) + (40)
ϱθ˙c(cosθcsinθcos(ϕϕc)sinθccosθ)+ϱsinθcϕ˙csinθsin(ϕϕc)+italic-ϱsubscript˙𝜃𝑐subscript𝜃𝑐𝜃italic-ϕsubscriptitalic-ϕ𝑐subscript𝜃𝑐𝜃limit-fromitalic-ϱsubscript𝜃𝑐subscript˙italic-ϕ𝑐𝜃italic-ϕsubscriptitalic-ϕ𝑐\displaystyle\left.\varrho\dot{\theta}_{c}(\cos\theta_{c}\sin\theta\cos(\phi-% \phi_{c})-\sin\theta_{c}\cos\theta)+\varrho\sin\theta_{c}\dot{\phi}_{c}\sin% \theta\sin(\phi-\phi_{c})+\right.italic_ϱ over˙ start_ARG italic_θ end_ARG start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( roman_cos italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT roman_sin italic_θ roman_cos ( italic_ϕ - italic_ϕ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) - roman_sin italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT roman_cos italic_θ ) + italic_ϱ roman_sin italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT over˙ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT roman_sin italic_θ roman_sin ( italic_ϕ - italic_ϕ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) +
VgD(Wϱcosθcosθcϱsinθsinθccos(ϕϕc))]+χeD/lcVgγI(P)κP,\displaystyle\left.\frac{V_{g}}{D}(W-\varrho\cos\theta\cos\theta_{c}-\varrho% \sin\theta\sin\theta_{c}\cos(\phi-\phi_{c}))\right]_{+}\frac{\chi e^{-D/l_{c}}% }{V_{g}}\gamma I(P)-\kappa P,divide start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG start_ARG italic_D end_ARG ( italic_W - italic_ϱ roman_cos italic_θ roman_cos italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - italic_ϱ roman_sin italic_θ roman_sin italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT roman_cos ( italic_ϕ - italic_ϕ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) ) ] start_POSTSUBSCRIPT + end_POSTSUBSCRIPT divide start_ARG italic_χ italic_e start_POSTSUPERSCRIPT - italic_D / italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG italic_γ italic_I ( italic_P ) - italic_κ italic_P ,
D𝐷\displaystyle Ditalic_D =W2+ϱ22Wϱcosθcosθc2Wϱsinθsinθccos(ϕϕc),absentsuperscript𝑊2superscriptitalic-ϱ22𝑊italic-ϱ𝜃subscript𝜃𝑐2𝑊italic-ϱ𝜃subscript𝜃𝑐italic-ϕsubscriptitalic-ϕ𝑐\displaystyle=\sqrt{W^{2}+\varrho^{2}-2W\varrho\cos\theta\cos\theta_{c}-2W% \varrho\sin\theta\sin\theta_{c}\cos(\phi-\phi_{c})},= square-root start_ARG italic_W start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_ϱ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 2 italic_W italic_ϱ roman_cos italic_θ roman_cos italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - 2 italic_W italic_ϱ roman_sin italic_θ roman_sin italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT roman_cos ( italic_ϕ - italic_ϕ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) end_ARG , (41)
χ𝜒\displaystyle\chiitalic_χ =12(111+(rmD)2),I(P)=1eMM0(1P)M/M0.formulae-sequenceabsent12111superscriptsubscript𝑟𝑚𝐷2𝐼𝑃1superscript𝑒𝑀subscript𝑀01𝑃𝑀subscript𝑀0\displaystyle=\frac{1}{2}\left(1-\frac{1}{\sqrt{1+\left(\frac{r_{m}}{D}\right)% ^{2}}}\right),\;\;\;I(P)=\frac{1-e^{-\frac{M}{M_{0}}(1-P)}}{M/M_{0}}.= divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 1 - divide start_ARG 1 end_ARG start_ARG square-root start_ARG 1 + ( divide start_ARG italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT end_ARG start_ARG italic_D end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG ) , italic_I ( italic_P ) = divide start_ARG 1 - italic_e start_POSTSUPERSCRIPT - divide start_ARG italic_M end_ARG start_ARG italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG ( 1 - italic_P ) end_POSTSUPERSCRIPT end_ARG start_ARG italic_M / italic_M start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG . (42)

We focus on the orbiting motion of a centrosome in the xy𝑥𝑦xyitalic_x italic_y-plane (where θc=π/2subscript𝜃𝑐𝜋2\theta_{c}=\pi/2italic_θ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_π / 2) with a constant radius ϱitalic-ϱ\varrhoitalic_ϱ from the sphere center and a constant angular frequency ϕ˙c=ωsubscript˙italic-ϕ𝑐𝜔\dot{\phi}_{c}=\omegaover˙ start_ARG italic_ϕ end_ARG start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_ω. With the change of variable ϕ¯=ϕϕc=ϕωt¯italic-ϕitalic-ϕsubscriptitalic-ϕ𝑐italic-ϕ𝜔𝑡\bar{\phi}=\phi-\phi_{c}=\phi-\omega tover¯ start_ARG italic_ϕ end_ARG = italic_ϕ - italic_ϕ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = italic_ϕ - italic_ω italic_t and the assumption that P(θ,ϕ,t)=P(θ,ϕ¯)𝑃𝜃italic-ϕ𝑡𝑃𝜃¯italic-ϕP(\theta,\phi,t)=P(\theta,\bar{\phi})italic_P ( italic_θ , italic_ϕ , italic_t ) = italic_P ( italic_θ , over¯ start_ARG italic_ϕ end_ARG ), we obtain the following equations

00\displaystyle 0 =Mf0AP(θ,ϕ¯)D(ϱ+Wsinθcos(ϕ¯))W2sinθdθdϕ¯,ϱ=WI2I1,\displaystyle=\frac{Mf_{0}}{A}\int\frac{P(\theta,\bar{\phi})}{D}\left(-\varrho% +W\sin\theta\cos(\bar{\phi})\right)W^{2}\sin\theta d\theta d\bar{\phi},% \rightarrow\varrho=W\frac{I_{2}}{I_{1}},= divide start_ARG italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_A end_ARG ∫ divide start_ARG italic_P ( italic_θ , over¯ start_ARG italic_ϕ end_ARG ) end_ARG start_ARG italic_D end_ARG ( - italic_ϱ + italic_W roman_sin italic_θ roman_cos ( over¯ start_ARG italic_ϕ end_ARG ) ) italic_W start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_sin italic_θ italic_d italic_θ italic_d over¯ start_ARG italic_ϕ end_ARG , → italic_ϱ = italic_W divide start_ARG italic_I start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG italic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG , (43)
ηϱω𝜂italic-ϱ𝜔\displaystyle\eta\varrho\omegaitalic_η italic_ϱ italic_ω =Mf0AP(θ,ϕ¯)Dsinθsin(ϕ¯)W3sinθdθdϕ¯,ηϱω=Mf0AI3,\displaystyle=\frac{Mf_{0}}{A}\int\frac{P(\theta,\bar{\phi})}{D}\sin\theta\sin% (\bar{\phi})W^{3}\sin\theta d\theta d\bar{\phi},\rightarrow\eta\varrho\omega=% \frac{Mf_{0}}{A}I_{3},= divide start_ARG italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_A end_ARG ∫ divide start_ARG italic_P ( italic_θ , over¯ start_ARG italic_ϕ end_ARG ) end_ARG start_ARG italic_D end_ARG roman_sin italic_θ roman_sin ( over¯ start_ARG italic_ϕ end_ARG ) italic_W start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_sin italic_θ italic_d italic_θ italic_d over¯ start_ARG italic_ϕ end_ARG , → italic_η italic_ϱ italic_ω = divide start_ARG italic_M italic_f start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_A end_ARG italic_I start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT , (44)
ωP(θ,ϕ¯)ϕ¯𝜔𝑃𝜃¯italic-ϕ¯italic-ϕ\displaystyle-\omega\frac{\partial P\left(\theta,\bar{\phi}\right)}{\partial% \bar{\phi}}- italic_ω divide start_ARG ∂ italic_P ( italic_θ , over¯ start_ARG italic_ϕ end_ARG ) end_ARG start_ARG ∂ over¯ start_ARG italic_ϕ end_ARG end_ARG =[ϱωsinθsin(ϕ¯)+VgD(Wϱsinθcos(ϕ¯))]+χeD/lcVgγI(P)κP,absentsubscriptdelimited-[]italic-ϱ𝜔𝜃¯italic-ϕsubscript𝑉𝑔𝐷𝑊italic-ϱ𝜃¯italic-ϕ𝜒superscript𝑒𝐷subscript𝑙𝑐subscript𝑉𝑔𝛾𝐼𝑃𝜅𝑃\displaystyle=\left[\varrho\omega\sin\theta\sin(\bar{\phi})+\frac{V_{g}}{D}(W-% \varrho\sin\theta\cos(\bar{\phi}))\right]_{+}\frac{\chi e^{-D/l_{c}}}{V_{g}}% \gamma I(P)-\kappa P,= [ italic_ϱ italic_ω roman_sin italic_θ roman_sin ( over¯ start_ARG italic_ϕ end_ARG ) + divide start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG start_ARG italic_D end_ARG ( italic_W - italic_ϱ roman_sin italic_θ roman_cos ( over¯ start_ARG italic_ϕ end_ARG ) ) ] start_POSTSUBSCRIPT + end_POSTSUBSCRIPT divide start_ARG italic_χ italic_e start_POSTSUPERSCRIPT - italic_D / italic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_POSTSUPERSCRIPT end_ARG start_ARG italic_V start_POSTSUBSCRIPT italic_g end_POSTSUBSCRIPT end_ARG italic_γ italic_I ( italic_P ) - italic_κ italic_P , (45)
D𝐷\displaystyle Ditalic_D =W2+ϱ22Wϱsinθcos(ϕ¯),absentsuperscript𝑊2superscriptitalic-ϱ22𝑊italic-ϱ𝜃¯italic-ϕ\displaystyle=\sqrt{W^{2}+\varrho^{2}-2W\varrho\sin\theta\cos(\bar{\phi})},= square-root start_ARG italic_W start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_ϱ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 2 italic_W italic_ϱ roman_sin italic_θ roman_cos ( over¯ start_ARG italic_ϕ end_ARG ) end_ARG , (46)

where

I1=P(θ,ϕ¯)Dsinθdθdϕ¯,subscript𝐼1𝑃𝜃¯italic-ϕ𝐷𝜃𝑑𝜃𝑑¯italic-ϕI_{1}=\int\frac{P(\theta,\bar{\phi})}{D}\sin\theta d\theta d\bar{\phi},italic_I start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = ∫ divide start_ARG italic_P ( italic_θ , over¯ start_ARG italic_ϕ end_ARG ) end_ARG start_ARG italic_D end_ARG roman_sin italic_θ italic_d italic_θ italic_d over¯ start_ARG italic_ϕ end_ARG , (47)
I2=P(θ,ϕ¯)Dsin2θcos(ϕ¯)𝑑θ𝑑ϕ¯,subscript𝐼2𝑃𝜃¯italic-ϕ𝐷superscript2𝜃¯italic-ϕdifferential-d𝜃differential-d¯italic-ϕI_{2}=\int\frac{P(\theta,\bar{\phi})}{D}\sin^{2}\theta\cos(\bar{\phi})d\theta d% \bar{\phi},italic_I start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = ∫ divide start_ARG italic_P ( italic_θ , over¯ start_ARG italic_ϕ end_ARG ) end_ARG start_ARG italic_D end_ARG roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ roman_cos ( over¯ start_ARG italic_ϕ end_ARG ) italic_d italic_θ italic_d over¯ start_ARG italic_ϕ end_ARG , (48)
I3=P(θ,ϕ¯)DW3sin2θsin(ϕ¯)𝑑θ𝑑ϕ¯.subscript𝐼3𝑃𝜃¯italic-ϕ𝐷superscript𝑊3superscript2𝜃¯italic-ϕdifferential-d𝜃differential-d¯italic-ϕI_{3}=\int\frac{P(\theta,\bar{\phi})}{D}W^{3}\sin^{2}\theta\sin(\bar{\phi})d% \theta d\bar{\phi}.italic_I start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = ∫ divide start_ARG italic_P ( italic_θ , over¯ start_ARG italic_ϕ end_ARG ) end_ARG start_ARG italic_D end_ARG italic_W start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ roman_sin ( over¯ start_ARG italic_ϕ end_ARG ) italic_d italic_θ italic_d over¯ start_ARG italic_ϕ end_ARG . (49)

Eq. (43) gives the radius of the orbit, and Eq. (44) gives the angular speed of the centrosome motion with the occupancy probability P𝑃Pitalic_P determined from Eq. (45). Altogether the system consists of two integrals and one differential equation for the occupancy probability P𝑃Pitalic_P.

This system of nonlinear integral-differential equations are solved as follows. For an initial guess of (ϱi,ωi)superscriptitalic-ϱ𝑖superscript𝜔𝑖(\varrho^{i},\omega^{i})( italic_ϱ start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT , italic_ω start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT ) we solve Eq. (45) to find the occupancy probability P𝑃Pitalic_P. We compute the integrals Issuperscript𝐼𝑠I^{\prime}sitalic_I start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT italic_s to update the value (ϱi+1,ωi+1)superscriptitalic-ϱ𝑖1superscript𝜔𝑖1(\varrho^{i+1},\omega^{i+1})( italic_ϱ start_POSTSUPERSCRIPT italic_i + 1 end_POSTSUPERSCRIPT , italic_ω start_POSTSUPERSCRIPT italic_i + 1 end_POSTSUPERSCRIPT ), and compute the loss function E=log(1+(δϱ)2+(δω)2)𝐸1superscript𝛿italic-ϱ2superscript𝛿𝜔2E=\log(1+(\delta\varrho)^{2}+(\delta\omega)^{2})italic_E = roman_log ( 1 + ( italic_δ italic_ϱ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( italic_δ italic_ω ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ), with δϱ=ϱi+1ϱi𝛿italic-ϱsuperscriptitalic-ϱ𝑖1superscriptitalic-ϱ𝑖\delta\varrho=\varrho^{i+1}-\varrho^{i}italic_δ italic_ϱ = italic_ϱ start_POSTSUPERSCRIPT italic_i + 1 end_POSTSUPERSCRIPT - italic_ϱ start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT, and δω=ωi+1ωi𝛿𝜔superscript𝜔𝑖1superscript𝜔𝑖\delta\omega=\omega^{i+1}-\omega^{i}italic_δ italic_ω = italic_ω start_POSTSUPERSCRIPT italic_i + 1 end_POSTSUPERSCRIPT - italic_ω start_POSTSUPERSCRIPT italic_i end_POSTSUPERSCRIPT. We update (ϱ,ω)italic-ϱ𝜔(\varrho,\omega)( italic_ϱ , italic_ω ) using an iterative gradient method for nonlinear systems to minimize the loss function within a set tolerance.

VI.5 Numerics for the linear stability of a centrosome in a spheroid

Here we present numerical details for the linear stability analysis of a centrosome in a general spheroid. To obtain high-order accuracy of the numerical linear stability analysis on this differential-integral system, we need to compute surface integral

Γf(𝐘)𝑑A,subscriptΓ𝑓𝐘differential-d𝐴\displaystyle\int_{\Gamma}f({\bf Y})dA,∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT italic_f ( bold_Y ) italic_d italic_A ,

with high-order spatial accuracy. We generate quadrature for spheroidal surfaces via a differential function φ𝜑\varphiitalic_φ, which maps ΓΓ\Gammaroman_Γ to the surface of the unit sphere (S2superscript𝑆2S^{2}italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT). We then transform a surface integral on ΓΓ\Gammaroman_Γ to one on the unit sphere as

Γ(𝐘)𝑑A(𝐘)=S2(φ(𝐗))|det(Jφ(𝐗))Jφt(𝐗)n^(𝐗)|𝑑A(𝐗),subscriptΓ𝐘differential-d𝐴𝐘subscriptsuperscript𝑆2𝜑𝐗subscript𝐽𝜑𝐗superscriptsubscript𝐽𝜑𝑡𝐗^𝑛𝐗differential-d𝐴𝐗\displaystyle\int_{\Gamma}\mathcal{F}({\bf Y})dA({\bf Y})=\int_{S^{2}}\mathcal% {F}(\varphi({\bf X}))\big{\lvert}\det(J_{\varphi}({\bf X}))J_{\varphi}^{-t}({% \bf X})\hat{n}({\bf X})\big{\rvert}dA({\bf X}),∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT caligraphic_F ( bold_Y ) italic_d italic_A ( bold_Y ) = ∫ start_POSTSUBSCRIPT italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_POSTSUBSCRIPT caligraphic_F ( italic_φ ( bold_X ) ) | roman_det ( italic_J start_POSTSUBSCRIPT italic_φ end_POSTSUBSCRIPT ( bold_X ) ) italic_J start_POSTSUBSCRIPT italic_φ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - italic_t end_POSTSUPERSCRIPT ( bold_X ) over^ start_ARG italic_n end_ARG ( bold_X ) | italic_d italic_A ( bold_X ) , (50)

where Jφsubscript𝐽𝜑J_{\varphi}italic_J start_POSTSUBSCRIPT italic_φ end_POSTSUBSCRIPT is the Jacobian and n^^𝑛\hat{n}over^ start_ARG italic_n end_ARG is the outward normal of the sphere. This transformation allows us to efficiently compute the force on the centrosome by spatially discretizing the surface using a quadrature rule for the sphere. In particular, we put three non-overlapping patches, two caps and a ring, described by

𝒜1subscript𝒜1\displaystyle\mathcal{A}_{1}caligraphic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ={𝐗S2|cos1(𝐗z^)π4},absentconditional-set𝐗superscript𝑆2superscript1𝐗^𝑧𝜋4\displaystyle=\left\{{\bf X}\in S^{2}\,|\,\cos^{-1}({\bf X}\cdot\hat{z})\leq% \frac{\pi}{4}\right\},= { bold_X ∈ italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | roman_cos start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( bold_X ⋅ over^ start_ARG italic_z end_ARG ) ≤ divide start_ARG italic_π end_ARG start_ARG 4 end_ARG } , (51)
𝒜2subscript𝒜2\displaystyle\mathcal{A}_{2}caligraphic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ={𝐗S2|π4cos1(𝐗z^)3π4},absentconditional-set𝐗superscript𝑆2𝜋4superscript1𝐗^𝑧3𝜋4\displaystyle=\left\{{\bf X}\in S^{2}\,|\,\frac{\pi}{4}\leq\cos^{-1}({\bf X}% \cdot\hat{z})\leq\frac{3\pi}{4}\right\},= { bold_X ∈ italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | divide start_ARG italic_π end_ARG start_ARG 4 end_ARG ≤ roman_cos start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( bold_X ⋅ over^ start_ARG italic_z end_ARG ) ≤ divide start_ARG 3 italic_π end_ARG start_ARG 4 end_ARG } , (52)
𝒜3subscript𝒜3\displaystyle\mathcal{A}_{3}caligraphic_A start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ={𝐗S2|cos1(𝐗z^)3π4}.absentconditional-set𝐗superscript𝑆2superscript1𝐗^𝑧3𝜋4\displaystyle=\left\{{\bf X}\in S^{2}\,|\,\cos^{-1}({\bf X}\cdot\hat{z})\geq% \frac{3\pi}{4}\right\}.= { bold_X ∈ italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT | roman_cos start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( bold_X ⋅ over^ start_ARG italic_z end_ARG ) ≥ divide start_ARG 3 italic_π end_ARG start_ARG 4 end_ARG } . (53)

Patches 𝒜1subscript𝒜1\mathcal{A}_{1}caligraphic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and 𝒜3subscript𝒜3\mathcal{A}_{3}caligraphic_A start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT can be discretize as a Disk, noting that sin(φ)φ𝜑𝜑\frac{\sin(\varphi)}{\varphi}divide start_ARG roman_sin ( italic_φ ) end_ARG start_ARG italic_φ end_ARG is smooth, hence, for the integral 0φ0f(φ)sinφφφ𝑑φsuperscriptsubscript0subscript𝜑0𝑓𝜑𝜑𝜑𝜑differential-d𝜑\int_{0}^{\varphi_{0}}f(\varphi)\frac{\sin{\varphi}}{\varphi}\varphi d\varphi∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUPERSCRIPT italic_f ( italic_φ ) divide start_ARG roman_sin italic_φ end_ARG start_ARG italic_φ end_ARG italic_φ italic_d italic_φ we can use the same quadrature rules used for the Disk [60], where φ𝜑\varphiitalic_φ takes the role of the radius.

Refer to caption
Figure 9: An illustration of patched quadrature. Yellow is for the ring patch 𝒜2subscript𝒜2{\cal A}_{2}caligraphic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, and blue (pink) is for the top (bottom) cap patch 𝒜1subscript𝒜1{\cal A}_{1}caligraphic_A start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT (𝒜3subscript𝒜3{\cal A}_{3}caligraphic_A start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT). Note that there is an accumulation of quadrature points around the union of the patches.

For 𝒜2subscript𝒜2\mathcal{A}_{2}caligraphic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT in the periodic direction a trapezoidal rule is used, on the non-periodic one a Legendre quadrature rule is used. For 𝒜1,3subscript𝒜13\mathcal{A}_{1,3}caligraphic_A start_POSTSUBSCRIPT 1 , 3 end_POSTSUBSCRIPT we use 10×10101010\times 1010 × 10 quadrature points, for 𝒜2subscript𝒜2\mathcal{A}_{2}caligraphic_A start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT we use 15×10151015\times 1015 × 10 quadrature points, periodic and height direction respectively. Fig. (9) is an example of the three non-overlapping patches of quadrature.

VI.6 Geometric intuition of two solutions

Here we provide an intuitive argument for no more than two eigenvalues in Eq. 22. We start with a similar equation

αx=β+11γx+t𝑑t=β+γlog(x+1x1)𝛼𝑥𝛽superscriptsubscript11𝛾𝑥𝑡differential-d𝑡𝛽𝛾𝑥1𝑥1\displaystyle\alpha x=-\beta+\int_{-1}^{1}\frac{\gamma}{x+t}dt=-\beta+\gamma% \log\left(\frac{x+1}{x-1}\right)italic_α italic_x = - italic_β + ∫ start_POSTSUBSCRIPT - 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 1 end_POSTSUPERSCRIPT divide start_ARG italic_γ end_ARG start_ARG italic_x + italic_t end_ARG italic_d italic_t = - italic_β + italic_γ roman_log ( divide start_ARG italic_x + 1 end_ARG start_ARG italic_x - 1 end_ARG ) (54)

This problem can be reduced to ask the number of solution of equation

ax+b=log(x+1x1).𝑎𝑥𝑏𝑥1𝑥1\displaystyle ax+b=\log\left(\frac{x+1}{x-1}\right).italic_a italic_x + italic_b = roman_log ( divide start_ARG italic_x + 1 end_ARG start_ARG italic_x - 1 end_ARG ) . (55)

From simple analysis, we can see that log(x+1x1)𝑥1𝑥1\log\left(\frac{x+1}{x-1}\right)roman_log ( divide start_ARG italic_x + 1 end_ARG start_ARG italic_x - 1 end_ARG ) has two vertical and one horizontal asymptotes. As we can see in Figure (10) there are only two cases when there are not two real solutions. Two of them are the degenerate case, where we have that the curve is tangent to the equation of the line and the other is when the line goes through exactly between the vertical asymptotes.

Refer to caption
Figure 10: The blue lines are the right-hand side of Eq. (55), and the dash lines are their two vertical asymptotes. The red, black and green lines are the left-hand side of Eq. (55), giving rise to three types of roots that Eq. (55) can have: Two roots at the intersections between the red line and the blue curve (in either the first or the third quadrant), two roots at the intersections between the black line and the blue curves, or no roots when the green curve crosses in between the asymptotes.

VI.7 Stable singular spectrum

In the main text we left out some interesting characteristics of the stable spectrum. It turns out that it has a diverse complicated structure for the slightest perturbation of the sphere, i.e. ratio α=1±ϵ𝛼plus-or-minus1italic-ϵ\alpha=1\pm\epsilonitalic_α = 1 ± italic_ϵ with 0<ϵ10italic-ϵmuch-less-than10<\epsilon\ll 10 < italic_ϵ ≪ 1. We start this discussion for the case when there is no centrosome dynamics, corresponding to 𝐯=𝟎𝐯0\bf v=0bold_v = bold_0 in Eqs. (19,20) and the eigenvalue equation is reduced to

Γp(𝐘)𝐘^𝑑A𝐘subscriptΓ𝑝𝐘^𝐘differential-dsubscript𝐴𝐘\displaystyle\int_{\Gamma}p({\bf Y})\hat{{\bf Y}}dA_{\bf Y}∫ start_POSTSUBSCRIPT roman_Γ end_POSTSUBSCRIPT italic_p ( bold_Y ) over^ start_ARG bold_Y end_ARG italic_d italic_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT =0absent0\displaystyle=0= 0 (56)
(σ+Ω~0(𝐘))p(𝐘)𝜎subscript~Ω0𝐘𝑝𝐘\displaystyle(\sigma+\tilde{\Omega}_{0}({\bf Y}))p({\bf Y})( italic_σ + over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) ) italic_p ( bold_Y ) =0.absent0\displaystyle=0.= 0 . (57)

Note that if p𝑝pitalic_p is a continuous function, the only solution to the system is p0𝑝0p\equiv 0italic_p ≡ 0. To find the rest of the spectrum we need to consider tempered distributions. To understand this eigenfunction, we see the following simpler eigenfunction system, given by

λf(x)=(x+ϵ)f(x)x[1,1]formulae-sequence𝜆𝑓𝑥𝑥italic-ϵ𝑓𝑥𝑥11\displaystyle\lambda f(x)=(x+\epsilon)f(x)\quad x\in[-1,1]italic_λ italic_f ( italic_x ) = ( italic_x + italic_ϵ ) italic_f ( italic_x ) italic_x ∈ [ - 1 , 1 ] (58)

The solutions to this problem is given by the pair (δ(xx0),x0)𝛿𝑥subscript𝑥0subscript𝑥0(\delta(x-x_{0}),x_{0})( italic_δ ( italic_x - italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) , italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ), where x0[1,1]subscript𝑥011x_{0}\in[-1,1]italic_x start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ∈ [ - 1 , 1 ]  [61]. We can generalize this solution by thinking in cylindrical coordinates. To start, remember that all the level sets 𝒜c={𝐘Γ|Ω~0(𝐘)=c}subscript𝒜𝑐conditional-set𝐘Γsubscript~Ω0𝐘𝑐\mathcal{A}_{c}=\{{\bf Y}\in\Gamma|\;\tilde{\Omega}_{0}({\bf Y})=-c\}caligraphic_A start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = { bold_Y ∈ roman_Γ | over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) = - italic_c } from axial symmetry are given by symmetric off center circles, which leads us to define zc0subscript𝑧𝑐0z_{c}\geq 0italic_z start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≥ 0, such that Ω~0((0,0,zc))=csubscript~Ω000subscript𝑧𝑐𝑐\tilde{\Omega}_{0}((0,0,z_{c}))=-cover~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( ( 0 , 0 , italic_z start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) ) = - italic_c. Using these, we can conclude that if there is a singular eigenfunction of the system, its eigenvalues need to be σ[max𝐘ΓΩ~0(𝐘)),min𝐘ΓΩ~0(𝐘))]\sigma\in[-\max_{{\bf Y}\in\Gamma}\tilde{\Omega}_{0}({\bf Y})),-\min_{{\bf Y}% \in\Gamma}\tilde{\Omega}_{0}({\bf Y}))]italic_σ ∈ [ - roman_max start_POSTSUBSCRIPT bold_Y ∈ roman_Γ end_POSTSUBSCRIPT over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) ) , - roman_min start_POSTSUBSCRIPT bold_Y ∈ roman_Γ end_POSTSUBSCRIPT over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) ) ] and its eigenfunction, pσsubscript𝑝𝜎p_{\sigma}italic_p start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT, has sing supp(pσ)𝒜σsing suppsubscript𝑝𝜎subscript𝒜𝜎\text{sing supp}(p_{\sigma})\subset\mathcal{A}_{\sigma}sing supp ( italic_p start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT ) ⊂ caligraphic_A start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT. With this idea in mind we can determine the eigenvalue spectrum {σ}=[max𝐘ΓΩ~0(𝐘)),min𝐘ΓΩ~0(𝐘))]\{\sigma\}=[-\max_{{\bf Y}\in\Gamma}\tilde{\Omega}_{0}({\bf Y})),-\min_{{\bf Y% }\in\Gamma}\tilde{\Omega}_{0}({\bf Y}))]{ italic_σ } = [ - roman_max start_POSTSUBSCRIPT bold_Y ∈ roman_Γ end_POSTSUBSCRIPT over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) ) , - roman_min start_POSTSUBSCRIPT bold_Y ∈ roman_Γ end_POSTSUBSCRIPT over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) ) ], with the corresponding generalized eigenfunctions

pσn,±(z,θ)={einθδ(z±zc)n,|n|2e±iθ(δ(zzc)δ(z+zc))n=1δ(zzc)+δ(z+zc)n=0superscriptsubscript𝑝𝜎𝑛plus-or-minus𝑧𝜃casesformulae-sequencesuperscript𝑒𝑖𝑛𝜃𝛿plus-or-minus𝑧subscript𝑧𝑐𝑛𝑛2otherwisesuperscript𝑒plus-or-minus𝑖𝜃𝛿𝑧subscript𝑧𝑐𝛿𝑧subscript𝑧𝑐𝑛1otherwise𝛿𝑧subscript𝑧𝑐𝛿𝑧subscript𝑧𝑐𝑛0otherwise\displaystyle p_{\sigma}^{n,\pm}(z,\theta)=\begin{cases}e^{in\theta}\delta(z% \pm z_{c})\quad n\in\mathbb{Z},\,|n|\geq 2\\ e^{\pm i\theta}(\delta(z-z_{c})-\delta(z+z_{c}))\quad n=1\\ \delta(z-z_{c})+\delta(z+z_{c})\quad n=0\end{cases}italic_p start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n , ± end_POSTSUPERSCRIPT ( italic_z , italic_θ ) = { start_ROW start_CELL italic_e start_POSTSUPERSCRIPT italic_i italic_n italic_θ end_POSTSUPERSCRIPT italic_δ ( italic_z ± italic_z start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) italic_n ∈ blackboard_Z , | italic_n | ≥ 2 end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL italic_e start_POSTSUPERSCRIPT ± italic_i italic_θ end_POSTSUPERSCRIPT ( italic_δ ( italic_z - italic_z start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) - italic_δ ( italic_z + italic_z start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) ) italic_n = 1 end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL italic_δ ( italic_z - italic_z start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) + italic_δ ( italic_z + italic_z start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ) italic_n = 0 end_CELL start_CELL end_CELL end_ROW (59)
Refer to caption
Figure 11: Graphical representation of pσn,±(z,θ)superscriptsubscript𝑝𝜎𝑛plus-or-minus𝑧𝜃p_{\sigma}^{n,\pm}(z,\theta)italic_p start_POSTSUBSCRIPT italic_σ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n , ± end_POSTSUPERSCRIPT ( italic_z , italic_θ ) for a spheroid with α=1.5𝛼1.5\alpha=1.5italic_α = 1.5. The curves over the surface represent the eigenfunction that is either 0 or infinity times a trigonometric function, as in Eq. (59). The red sphere denotes the stably centered aster.

A schematic of the generalized eigenfunctions is shown in Fig. (11). We remark how much the structure of the eigenvectors change with a perturbation of the sphere. For the sphere, there was only one eigenvalue associated with modes that do not generate centrosome motion. Its corresponding eigenspace could be decomposed into a countable set of spherical harmonics. Meanwhile, for the ellipsoidal case, we have a continuum spectrum of modes that do not generate centrosome motion, where each eigenspace is spanned by a countable set of generalized functions.

Incredibly, we find more structure of the solution when we depart from the standard analysis and reinterpret the integrals in Eq. (24) as principal value integrals, opening the possibility to find new nonlinear solutions of the equation in the range [max𝐘ΓΩ~0(𝐘)),min𝐘ΓΩ~0(𝐘)][-\max_{{\bf Y}\in\Gamma}\tilde{\Omega}_{0}({\bf Y})),-\min_{{\bf Y}\in\Gamma}% \tilde{\Omega}_{0}({\bf Y})][ - roman_max start_POSTSUBSCRIPT bold_Y ∈ roman_Γ end_POSTSUBSCRIPT over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) ) , - roman_min start_POSTSUBSCRIPT bold_Y ∈ roman_Γ end_POSTSUBSCRIPT over~ start_ARG roman_Ω end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( bold_Y ) ]. Numerically we can find there is at least one solution for each index on the parameter range that we studied. That being said, the corresponding eigenvectors are decaying modes and take singular values, hence not of physical interest.

VI.8 Probabilistic derivation of the overlap model

Here we provide detailed derivations of the exponential model for the interaction between force generators and the astral MTs. We present two different approaches. One is a direct, explicit derivation with all the hypothesis of the problem and substantial algebraic details, and the other simpler calculation making an analogy to picking balls of different colors, with less details. To simplify the notation, we introduce the following events

  • 𝐀𝐘=subscript𝐀𝐘absent{\bf A}_{\bf Y}=bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT = MT hits CFGY (CFG centered at point 𝐘𝐘{\bf Y}bold_Y)

  • 𝐁𝐘=subscript𝐁𝐘absent{\bf B}_{\bf Y}=bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT = MT binds to CFGY

  • 𝐂𝐘=subscript𝐂𝐘absent{\bf C}_{\bf Y}=bold_C start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT = CFGY is occupied

  • 𝐃𝐘n=subscriptsuperscript𝐃𝑛𝐘absent{\bf D}^{n}_{\bf Y}=bold_D start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT = There are n𝑛nitalic_n other CFG competing for the same MT that already binds to a CFGY

Note that in this notation P(𝐘)=(𝐂𝐘)𝑃𝐘subscript𝐂𝐘P({\bf Y})=\mathbb{P}({\bf C}_{\bf Y})italic_P ( bold_Y ) = blackboard_P ( bold_C start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT ).

Let U𝑈Uitalic_U be a patch over the surface ΓΓ\Gammaroman_Γ, centered at 𝐘𝐘\bf Ybold_Y, of area AUsubscript𝐴𝑈A_{U}italic_A start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT such that 𝐗Ufor-all𝐗𝑈\forall{\bf X}\in U∀ bold_X ∈ italic_U, |P(𝐗)P(𝐘)|<<1much-less-than𝑃𝐗𝑃𝐘1|P({\bf X})-P({\bf Y})|<<1| italic_P ( bold_X ) - italic_P ( bold_Y ) | < < 1. In the patch the total number of force generators is given by NU=MUρ(𝐗)𝑑AMAUρ(𝐘)subscript𝑁𝑈𝑀subscript𝑈𝜌𝐗differential-d𝐴𝑀subscript𝐴𝑈𝜌𝐘N_{U}=M\int_{U}\rho({\bf X})dA\approx MA_{U}\rho({\bf Y})italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT = italic_M ∫ start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT italic_ρ ( bold_X ) italic_d italic_A ≈ italic_M italic_A start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT italic_ρ ( bold_Y ). To start the derivation we see that the change in P(𝐘)𝑃𝐘P({\bf Y})italic_P ( bold_Y ) in a time-step ΔtΔ𝑡\Delta troman_Δ italic_t is given by

dP(𝐘)=κP(𝐘)Δt+(MT binds to CFGY in Δt).𝑑𝑃𝐘𝜅𝑃𝐘Δ𝑡MT binds to CFGY in Δt\displaystyle dP({\bf Y})=-\kappa P({\bf Y})\Delta t+\mathbb{P}(\text{MT binds% to CFG${}_{\bf Y}$ in $\Delta t$}).italic_d italic_P ( bold_Y ) = - italic_κ italic_P ( bold_Y ) roman_Δ italic_t + blackboard_P ( MT binds to CFG in roman_Δ italic_t ) . (60)

By the law of total probabilities, we have

(MT binds to CFGY in Δt)=Ω(𝐘)(𝐁𝐘|𝐀𝐘)Δt.MT binds to CFGY in ΔtΩ𝐘conditionalsubscript𝐁𝐘subscript𝐀𝐘Δ𝑡\displaystyle\mathbb{P}(\text{MT binds to CFG${}_{\bf Y}$ in $\Delta t$})=% \Omega({\bf Y})\,\mathbb{P}({\bf B}_{\bf Y}|{\bf A}_{\bf Y})\,\Delta t.blackboard_P ( MT binds to CFG in roman_Δ italic_t ) = roman_Ω ( bold_Y ) blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT ) roman_Δ italic_t . (61)

The details of modeling CFG will be expressed through the term (𝐁𝐘|𝐀𝐘)conditionalsubscript𝐁𝐘subscript𝐀𝐘\mathbb{P}({\bf B}_{\bf Y}|{\bf A}_{\bf Y})blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT ). For stochiometric interactions, i.e., a free astral MT cannot bind to a CFG that is already occupied, we have (𝐁𝐘|𝐂𝐘)=0conditionalsubscript𝐁𝐘subscript𝐂𝐘0\mathbb{P}({\bf B}_{\bf Y}|{\bf C}_{\bf Y})=0blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_C start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT ) = 0. Therefore, by the law of total probabilities, we have

(𝐁𝐘|𝐀𝐘)=(1P(𝐘))(𝐁𝐘|𝐀𝐘,𝐂𝐘c)conditionalsubscript𝐁𝐘subscript𝐀𝐘1𝑃𝐘conditionalsubscript𝐁𝐘subscript𝐀𝐘superscriptsubscript𝐂𝐘𝑐\displaystyle\mathbb{P}({\bf B}_{\bf Y}|{\bf A}_{\bf Y})=(1-P({\bf Y}))\mathbb% {P}({\bf B}_{\bf Y}|{\bf A}_{\bf Y},{\bf C}_{\bf Y}^{c})blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT ) = ( 1 - italic_P ( bold_Y ) ) blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT , bold_C start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT ) (62)

We first note that if there was no interaction with other CFGs, then (𝐁𝐘|𝐂𝐘c)=1conditionalsubscript𝐁𝐘superscriptsubscript𝐂𝐘𝑐1\mathbb{P}({\bf B}_{\bf Y}|{\bf C}_{\bf Y}^{c})=1blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_C start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_c end_POSTSUPERSCRIPT ) = 1, yielding the independent motor model. However in this work we consider situations where there can be overlap between CFGs. To be specific we consider that if a MT hits a point where there is an overlap between n𝑛nitalic_n unoccupied CFGs, this MT will bind randomly to only one of these. In terms of probabilities, it means that (𝐁𝐘|𝐀𝐘,𝐂𝐘,𝐃𝐘n)=11+nconditionalsubscript𝐁𝐘subscript𝐀𝐘subscript𝐂𝐘superscriptsubscript𝐃𝐘𝑛11𝑛\mathbb{P}({\bf B}_{\bf Y}|{\bf A}_{\bf Y},{\bf C}_{\bf Y},{\bf D}_{\bf Y}^{n}% )=\frac{1}{1+n}blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT , bold_C start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT , bold_D start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ) = divide start_ARG 1 end_ARG start_ARG 1 + italic_n end_ARG. Now, for a CFGX to be available to compete for the MT it needs to be unbound and there has to be overlap between CFGX and CFGY. Since these two events are independent, we just need to approximate the probability that there is overlap between different CFGs in the patch U𝑈Uitalic_U. For this we do a simple homogeneity approximation. From here we can see that D𝐘nsuperscriptsubscript𝐷𝐘𝑛D_{\bf Y}^{n}italic_D start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT is a binomial distribution (in n𝑛nitalic_n) of parameters (NU1,aAu(1P(Y)))subscript𝑁𝑈1𝑎subscript𝐴𝑢1𝑃𝑌(N_{U}-1,\frac{a}{A_{u}}(1-P(Y)))( italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT - 1 , divide start_ARG italic_a end_ARG start_ARG italic_A start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT end_ARG ( 1 - italic_P ( italic_Y ) ) ), where a=πrm2𝑎𝜋superscriptsubscript𝑟𝑚2a=\pi r_{m}^{2}italic_a = italic_π italic_r start_POSTSUBSCRIPT italic_m end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the discal area of the CFG binding domain. In other words we have

(𝐃𝐘n|𝐀𝐘)=(NU1n)((1P)aAu)n(1(1P)aAu)NU1nconditionalsubscriptsuperscript𝐃𝑛𝐘subscript𝐀𝐘binomialsubscript𝑁𝑈1𝑛superscript1𝑃𝑎subscript𝐴𝑢𝑛superscript11𝑃𝑎subscript𝐴𝑢subscript𝑁𝑈1𝑛\mathbb{P}({\bf D}^{n}_{\bf Y}|{\bf A}_{\bf Y})={N_{U}-1\choose n}\left((1-P)% \frac{a}{A_{u}}\right)^{n}\left(1-(1-P)\frac{a}{A_{u}}\right)^{N_{U}-1-n}blackboard_P ( bold_D start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT ) = ( binomial start_ARG italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT - 1 end_ARG start_ARG italic_n end_ARG ) ( ( 1 - italic_P ) divide start_ARG italic_a end_ARG start_ARG italic_A start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ( 1 - ( 1 - italic_P ) divide start_ARG italic_a end_ARG start_ARG italic_A start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT - 1 - italic_n end_POSTSUPERSCRIPT

Putting these together using the law of total probabilities, we then have to evaluate the sum

(𝐁𝐘|𝐀𝐘)=(1P(𝐘))n=0NU111+n(NU1n)((1P)aAu)n(1(1P)aAu)NU1n.conditionalsubscript𝐁𝐘subscript𝐀𝐘1𝑃𝐘superscriptsubscript𝑛0subscript𝑁𝑈111𝑛binomialsubscript𝑁𝑈1𝑛superscript1𝑃𝑎subscript𝐴𝑢𝑛superscript11𝑃𝑎subscript𝐴𝑢subscript𝑁𝑈1𝑛\displaystyle\mathbb{P}({\bf B}_{\bf Y}|{\bf A}_{\bf Y})=(1-P({\bf Y}))\sum_{n% =0}^{N_{U}-1}\frac{1}{1+n}{N_{U}-1\choose n}\left((1-P)\frac{a}{A_{u}}\right)^% {n}\left(1-(1-P)\frac{a}{A_{u}}\right)^{N_{U}-1-n}.blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT ) = ( 1 - italic_P ( bold_Y ) ) ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT - 1 end_POSTSUPERSCRIPT divide start_ARG 1 end_ARG start_ARG 1 + italic_n end_ARG ( binomial start_ARG italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT - 1 end_ARG start_ARG italic_n end_ARG ) ( ( 1 - italic_P ) divide start_ARG italic_a end_ARG start_ARG italic_A start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ( 1 - ( 1 - italic_P ) divide start_ARG italic_a end_ARG start_ARG italic_A start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT - 1 - italic_n end_POSTSUPERSCRIPT . (63)

This sum can be easily evaluated by first letting q=(1P(𝐘)aAU)𝑞1𝑃𝐘𝑎subscript𝐴𝑈q=(1-P(\mathbf{Y})\frac{a}{A_{U}})italic_q = ( 1 - italic_P ( bold_Y ) divide start_ARG italic_a end_ARG start_ARG italic_A start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT end_ARG ) and J=NU1𝐽subscript𝑁𝑈1J=N_{U}-1italic_J = italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT - 1, then we have

n=0J(Jn)qnn+1(1q)Jn=1q0qn=0J(Jn)xn(1q)Jndx.superscriptsubscript𝑛0𝐽binomial𝐽𝑛superscript𝑞𝑛𝑛1superscript1𝑞𝐽𝑛1𝑞superscriptsubscript0𝑞superscriptsubscript𝑛0𝐽binomial𝐽𝑛superscript𝑥𝑛superscript1𝑞𝐽𝑛𝑑𝑥\displaystyle\sum_{n=0}^{J}{J\choose n}\frac{q^{n}}{n+1}(1-q)^{J-n}=\frac{1}{q% }\int_{0}^{q}\sum_{n=0}^{J}{J\choose n}x^{n}(1-q)^{J-n}dx.∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT ( binomial start_ARG italic_J end_ARG start_ARG italic_n end_ARG ) divide start_ARG italic_q start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT end_ARG start_ARG italic_n + 1 end_ARG ( 1 - italic_q ) start_POSTSUPERSCRIPT italic_J - italic_n end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG italic_q end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_q end_POSTSUPERSCRIPT ∑ start_POSTSUBSCRIPT italic_n = 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_J end_POSTSUPERSCRIPT ( binomial start_ARG italic_J end_ARG start_ARG italic_n end_ARG ) italic_x start_POSTSUPERSCRIPT italic_n end_POSTSUPERSCRIPT ( 1 - italic_q ) start_POSTSUPERSCRIPT italic_J - italic_n end_POSTSUPERSCRIPT italic_d italic_x . (64)

Using Newton’s binomial theorem and integrating, we obtain the final expression

(𝐁𝐘|𝐀𝐘)=1NUaAu(1(1(1P)aAu)NU.)\displaystyle\mathbb{P}({\bf B}_{\bf Y}|{\bf A}_{\bf Y})=\frac{1}{N_{U}\frac{a% }{A_{u}}}\left(1-(1-(1-P)\frac{a}{A_{u}})^{N_{U}}.\right)blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT ) = divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT divide start_ARG italic_a end_ARG start_ARG italic_A start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT end_ARG end_ARG ( 1 - ( 1 - ( 1 - italic_P ) divide start_ARG italic_a end_ARG start_ARG italic_A start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT end_POSTSUPERSCRIPT . ) (65)

From this expression, we obtain the following three interesting results.

  1. 1.

    If there is no overlap between CFGs, Au=asubscript𝐴𝑢𝑎A_{u}=aitalic_A start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT = italic_a and NU=1subscript𝑁𝑈1N_{U}=1italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT = 1, and we recover the independent CFG model where

    (𝐁𝐘|𝐀𝐘)=1Pconditionalsubscript𝐁𝐘subscript𝐀𝐘1𝑃\displaystyle\mathbb{P}({\bf B}_{\bf Y}|{\bf A}_{\bf Y})=1-Pblackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT ) = 1 - italic_P (66)
  2. 2.

    Taking the limit when aAu0𝑎subscript𝐴𝑢0\frac{a}{A_{u}}\to 0divide start_ARG italic_a end_ARG start_ARG italic_A start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT end_ARG → 0 we obtain the interaction probability (P)𝑃{\cal I}(P)caligraphic_I ( italic_P ) used in this work

    (𝐁𝐘|𝐀𝐘)=1eaMρ(1P)aMρ=(P).conditionalsubscript𝐁𝐘subscript𝐀𝐘1superscript𝑒𝑎𝑀𝜌1𝑃𝑎𝑀𝜌𝑃\displaystyle\mathbb{P}({\bf B}_{\bf Y}|{\bf A}_{\bf Y})=\frac{1-e^{-aM\rho(1-% P)}}{aM\rho}={\cal I}(P).blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT ) = divide start_ARG 1 - italic_e start_POSTSUPERSCRIPT - italic_a italic_M italic_ρ ( 1 - italic_P ) end_POSTSUPERSCRIPT end_ARG start_ARG italic_a italic_M italic_ρ end_ARG = caligraphic_I ( italic_P ) . (67)
  3. 3.

    If the CFGs are perfectly stacked one over each other (instead of being randomly distributed), this means that NUsubscript𝑁𝑈N_{U}italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT is independent of a𝑎aitalic_a, so we can just call it N𝑁Nitalic_N and we get

    (𝐁𝐘|𝐀𝐘)=1PNNconditionalsubscript𝐁𝐘subscript𝐀𝐘1superscript𝑃𝑁𝑁\displaystyle\mathbb{P}({\bf B}_{\bf Y}|{\bf A}_{\bf Y})=\frac{1-P^{N}}{N}blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT ) = divide start_ARG 1 - italic_P start_POSTSUPERSCRIPT italic_N end_POSTSUPERSCRIPT end_ARG start_ARG italic_N end_ARG (68)

Below we present a second derivation, simpler in the algebra, that hides many of the details of the interactions. First we define the event

𝐄U=MT hits patch U,subscript𝐄𝑈MT hits patch 𝑈{\bf E}_{U}=\text{MT hits patch }U,bold_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT = MT hits patch italic_U ,

using law of total probabilities we have

(𝐁𝐘|𝐄U)=(𝐁𝐘|𝐄U,𝐀𝐘)(𝐀𝐘|𝐄U)+(𝐁𝐘|𝐄U,𝐀𝐘C)(𝐀𝐘C|𝐄U).conditionalsubscript𝐁𝐘subscript𝐄𝑈conditionalsubscript𝐁𝐘subscript𝐄𝑈subscript𝐀𝐘conditionalsubscript𝐀𝐘subscript𝐄𝑈conditionalsubscript𝐁𝐘subscript𝐄𝑈superscriptsubscript𝐀𝐘𝐶conditionalsuperscriptsubscript𝐀𝐘𝐶subscript𝐄𝑈\displaystyle\mathbb{P}({\bf B}_{\bf Y}|{\bf E}_{U})=\mathbb{P}({\bf B}_{\bf Y% }|{\bf E}_{U},{\bf A}_{\bf Y})\mathbb{P}({\bf A}_{\bf Y}|{\bf E}_{U})+\mathbb{% P}({\bf B}_{\bf Y}|{\bf E}_{U},{\bf A}_{\bf Y}^{C})\mathbb{P}({\bf A}_{\bf Y}^% {C}|{\bf E}_{U}).blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ) = blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT , bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT ) blackboard_P ( bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ) + blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT , bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT ) blackboard_P ( bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_C end_POSTSUPERSCRIPT | bold_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ) . (69)

The probability of binding to CFGY with no MT hitting the area that it covers, is 0. Further, conditioning over 𝐄Usubscript𝐄𝑈{\bf E}_{U}bold_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT and 𝐀𝐘subscript𝐀𝐘{\bf A}_{\bf Y}bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT is redundant, since the latter implies the former. Thus the expression is simplified to

(𝐁𝐘|𝐄U)=(𝐁𝐘|𝐀𝐘)(𝐀𝐘|𝐄U).conditionalsubscript𝐁𝐘subscript𝐄𝑈conditionalsubscript𝐁𝐘subscript𝐀𝐘conditionalsubscript𝐀𝐘subscript𝐄𝑈\displaystyle\mathbb{P}({\bf B}_{\bf Y}|{\bf E}_{U})=\mathbb{P}({\bf B}_{\bf Y% }|{\bf A}_{\bf Y})\mathbb{P}({\bf A}_{\bf Y}|{\bf E}_{U}).blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ) = blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT ) blackboard_P ( bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ) . (70)

We are interested in (re)determining (𝐁𝐘|𝐀𝐘)conditionalsubscript𝐁𝐘subscript𝐀𝐘\mathbb{P}({\bf B}_{\bf Y}|{\bf A}_{\bf Y})blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT ), but (𝐁𝐘|𝐄U)conditionalsubscript𝐁𝐘subscript𝐄𝑈\mathbb{P}({\bf B}_{\bf Y}|{\bf E}_{U})blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ) has the advantage that every CFG within patch U𝑈Uitalic_U can be treated equally. To simplify the computation of this quantity, we set up an equivalent but simpler problem. Take NUsubscript𝑁𝑈N_{U}italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT balls, with one of them being red, and put them in two different boxes, each ball with probability (1P)aAu1𝑃𝑎subscript𝐴𝑢(1-P)\frac{a}{A_{u}}( 1 - italic_P ) divide start_ARG italic_a end_ARG start_ARG italic_A start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT end_ARG being in box one, and in the other box with the complement probability. Next, one ball is picked from box one (if there is any ball). We pose the question: What is the probability that the red ball is picked?

It is not difficult to see that getting the red ball has a probability of (𝐁𝐘|𝐄U)conditionalsubscript𝐁𝐘subscript𝐄𝑈\mathbb{P}({\bf B}_{\bf Y}|{\bf E}_{U})blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ). In this setting, we just need to determine the probability that there is at least one ball in box one and that the picked ball is red. Since all of the balls are the same, these two probabilities are independent, so it is straightforward that

(𝐁𝐘|𝐄U)=1NU(1(1(1P)aAu)NU).conditionalsubscript𝐁𝐘subscript𝐄𝑈1subscript𝑁𝑈1superscript11𝑃𝑎subscript𝐴𝑢subscript𝑁𝑈\mathbb{P}({\bf B}_{\bf Y}|{\bf E}_{U})=\frac{1}{N_{U}}\left(1-(1-(1-P)\frac{a% }{A_{u}})^{N_{U}}\right).blackboard_P ( bold_B start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ) = divide start_ARG 1 end_ARG start_ARG italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT end_ARG ( 1 - ( 1 - ( 1 - italic_P ) divide start_ARG italic_a end_ARG start_ARG italic_A start_POSTSUBSCRIPT italic_u end_POSTSUBSCRIPT end_ARG ) start_POSTSUPERSCRIPT italic_N start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT end_POSTSUPERSCRIPT ) .

Noting that (𝐀𝐘|𝐄U)=aAUconditionalsubscript𝐀𝐘subscript𝐄𝑈𝑎subscript𝐴𝑈\mathbb{P}({\bf A}_{\bf Y}|{\bf E}_{U})=\frac{a}{A_{U}}blackboard_P ( bold_A start_POSTSUBSCRIPT bold_Y end_POSTSUBSCRIPT | bold_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ) = divide start_ARG italic_a end_ARG start_ARG italic_A start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT end_ARG finalizes the proof.

VI.9 Aster dynamics inside a cuboid

Refer to caption
Figure 12: Distribution of pulling force on a centrosome in a square slab-like cell. (A) A square chamber with CFGs on the sidewall, similar to the microfabricated chamber with dynein motors attached on the sidewalls in [14]. (B)-(C) By computing the force required to fix the centrosome at a point inside the chamber, we obtain the direction fields of the pulling force on the centrosome inside a square chamber for characteristic MT length of lc=20μmsubscript𝑙𝑐20𝜇𝑚l_{c}=20\mu mitalic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 20 italic_μ italic_m and 7μm7𝜇𝑚7\mu m7 italic_μ italic_m, respectively. Red circles are for stable fixed points (where the net forces converge), yellow circles are for saddle fixed points (where the net forces converge in one direction and diverge in another), and orange circles denote unstable fixed points (where the net forces diverge).

We conducted simulations of the S-model with a centrosome placed inside a cuboid of dimensions 15×15×3μm15153𝜇𝑚15\times 15\times 3\;\mu m15 × 15 × 3 italic_μ italic_m with CFGs positioned exclusively on the peripheral sides (see Fig. 12A), imitating the microchamber in the experiments [14, 16, 15]. For a given characteristic MT length, we compute the pulling force distribution minus the force needed to keep the centrosome at a given point inside the chamber. When the average MT length (lc=20μmsubscript𝑙𝑐20𝜇𝑚l_{c}=20\;\mu mitalic_l start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 20 italic_μ italic_m) is comparable to the length of the microchamber, a singular fixed point emerges at the chamber’s center (Fig. 12B). For shorter MTs, a constellation of fixed points appears, distributed around the central region (Fig. 12C). The central fixed point becomes unstable, while eight additional fixed points formed, with those near the chamber’s corners exhibiting stability (red in Fig. 12C), and the intermediates acting as saddle points (yellow in Fig. 12C). Previous theoretical models successfully explained the positioning of the aster in these microchambers by a balance of pulling forces from motors and pushing forces from MT growth and friction with the chamber periphery [14, 16, 15]. Our findings provide an alternative explanation for the positioning of centrosomes in microchambers, suggesting it results from the combined effect of pulling forces and the stoichiometric interactions between MTs and molecular motors.