ABSTRACT Title of Dissertation: SIMULATING MEMBRANE-BOUND CYTOSKELETAL DYNAMICS Haoran Ni Doctor of Philosophy, 2023 Dissertation Directed by: Professor Garegin A. Papoian Department of Chemistry and Biochemistry The cell membrane defines the shape of the cell and plays an indispensable role in bridging intra- and extra-cellular environments. The membrane, consisting of a lipid bilayer and various attaching proteins, mechanochemically interacts with the active cytoskeletal network that dynam- ically self-organizes, playing a vital role in cellular biomechanics and mechanosensing. Compre- hensive simulations of membrane-cytoskeleton dynamics can bring insight in understanding how the cell mechanochemically responds to external signals, but a computational model that cap- tures the complex cytoskeleton-membrane with both refined details and computational efficiency is lacking. To address this, we introduce in this thesis a triangulated membrane model and in- corporate it with the active biological matter simulation platform MEDYAN (“Mechanochemical Dynamics of Active Networks”). This model accurately captures the membrane physical prop- erties, showing how the membrane rigidity, the structure of actin networks and local chemical environments regulate the membrane deformations. Then, we present a new method for simulat- ing membrane proteins, using stochastic reaction-diffusion sampling on unstructured membrane meshes. By incorporating a surface potential energy field into the reaction-diffusion sampling, we demonstrate rich membrane protein collective behaviors such as confined diffusion, liquid- liquid phase separation and membrane curvature sensing. Finally, in order to capture stretching, bending, shearing and twisting of actin filaments which are not all available with traditional acto- myosin simulations, we introduce new finite-radius filament models based off Cosserat theory of elastic rods, with efficient implementation using finite-dimensional configurational spaces. Us- ing the new filament models, we show that the filaments’ torsional compliance can induce chiral symmetry breaking in a cross-linked actin bundle. All the new models are implemented in the MEDYAN platform, shedding light on whole cell simulations, paving way for a better under- standing of the membrane-cytoskeleton system and its role in cellular dynamics. SIMULATING MEMBRANE-BOUND CYTOSKELETAL DYNAMICS by Haoran Ni Dissertation submitted to the Faculty of the Graduate School of the University of Maryland, College Park in partial fulfillment of the requirements for the degree of Doctor of Philosophy 2023 Advisory Committee: Professor Garegin A. Papoian, Chair/Advisor Professor Arpita Upadhyaya Professor Pratyush Tiwary Professor Sergei Sukharev Professor Jeffery B. Klauda © Copyright by Haoran Ni 2023 Acknowledgments I owe my gratitude to all the people who have made this thesis possible and because of whom my graduate experience in the past 6.5 years has been one that I will cherish forever. First and foremost I’d like to thank my advisor, Garegin Papoian. I have gained a lot by discussing with you about scientific, research and coding topics, and more importantly the way of prioritizing things as a researcher. I am grateful to be mentored by such an inspiring scientist. I would also like to thank my committee members Arpita Upadhyaya, Pratyush Tiwary, Sergei Sukharev and Jeffery Klauda. I appreciate all your great help in guiding this thesis and the opportunity to learn from you. I would extend my thanks to the Papoian lab members, who have been colleges and friends along the way. Haiqing Zhao, Hao Wu, Mary Pitman, Ryan Nival, as well as the MEDYAN de- velopers James Komianos, Aravind Chandrasekaran, Carlos Floyd, Qin Ni, Nathan Zimmerberg and Mengxin Gu have made the my lab experience joyful. It has been a great pleasure to talk about scientific problems, debug erroneous codes or chat about life in general. I have learned a lot from you and am grateful for all the helps I received, thank you. I would also like to acknowledge and thank the University of Maryland organizations I have been a member of, including the Biophysics program, the Institute for Physical Science and Technology and the Physics of Living Systems network. The faculty and students in these organizations have broaden my views with research in a wide range of disciplines and I am ii grateful for inspiring discussions with people outside my lab. I would also like to thank the staff members in these organizations for delivering excellent services. Finally, I owe my deepest thanks to my family. To my parents, Yongping Ni and Xunjuan Qiu, my grandparents and my cousins, thank you for being my constant source of strength, even you are overseas thousands of miles away. To all my friends along the way, thank you all. iii Table of Contents Acknowledgements ii Table of Contents iv List of Tables vii List of Figures viii List of Abbreviations x Chapter 1: Introduction 1 1.1 Biological membrane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2 1.2 Cytoskeleton . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 3 1.3 Cytoskeleton-membrane interactions . . . . . . . . . . . . . . . . . . . . . . . . 7 1.4 Outline of Thesis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10 Chapter 2: Simulating deformable vesicles containing complex cytoskeletal networks 12 2.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12 2.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 2.2.1 MEDYAN Simulations . . . . . . . . . . . . . . . . . . . . . . . . . . . 15 2.2.2 Membrane mechanics and discretization . . . . . . . . . . . . . . . . . . 15 2.2.3 Adaptive surface remeshing . . . . . . . . . . . . . . . . . . . . . . . . 25 2.2.4 Interactions between the membrane and the cytoskeleton . . . . . . . . . 26 2.2.5 Computational efficiency . . . . . . . . . . . . . . . . . . . . . . . . . . 29 2.2.6 Code availability . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 2.3 Results and Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30 2.4 Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36 Chapter 3: Simulating membrane proteins and their dynamics 39 3.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 39 3.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 41 3.2.1 Thermodynamics of interacting proteins on the surface . . . . . . . . . . 41 3.2.2 Surface discretization and stochastic diffusion sampling . . . . . . . . . . 43 3.2.3 Membrane-related reactions . . . . . . . . . . . . . . . . . . . . . . . . 46 3.2.4 Free energy of the particle system . . . . . . . . . . . . . . . . . . . . . 48 3.2.5 Membrane curvature sensing and generation model . . . . . . . . . . . . 49 3.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50 iv 3.3.1 Energy barriers acting as protein confinements decrease diffusivity . . . . 50 3.3.2 Pairwise protein interactions promote LLPS . . . . . . . . . . . . . . . . 53 3.3.3 Concentration sorting by local curvature . . . . . . . . . . . . . . . . . . 54 3.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 Chapter 4: Cosserat rod model of actin filaments 61 4.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61 4.2 Methods . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64 4.2.1 Cosserat theory of elastic rods . . . . . . . . . . . . . . . . . . . . . . . 66 4.2.2 Variational approach to rod mechanics . . . . . . . . . . . . . . . . . . . 69 4.2.3 Geodesic models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 71 4.2.4 MEDYAN model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73 4.2.5 Dynamical model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 74 4.2.6 Binding to surface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 75 4.3 Results . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 4.3.1 Model comparisons . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 77 4.3.2 Bundle study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 80 4.3.3 MEDYAN Implementation . . . . . . . . . . . . . . . . . . . . . . . . . 87 4.4 Discussion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 88 Chapter 5: Summary and future directions 93 Appendix A: Supporting information for Chapter 2 96 A.1 Simulation parameters and initial setups . . . . . . . . . . . . . . . . . . . . . . 96 A.1.1 Simulations for interactions between actin and membrane vesicle . . . . . 96 A.1.2 Perpendicular bundle protrusion against the membrane . . . . . . . . . . 97 A.1.3 Non-perpendicular bundle protrusion against the membrane . . . . . . . . 97 A.2 Representation of membrane surface using a triangular mesh . . . . . . . . . . . 98 A.3 Benchmark on membrane adaptive remeshing . . . . . . . . . . . . . . . . . . . 100 A.4 Membrane bending energy on discretized surface . . . . . . . . . . . . . . . . . 100 A.4.1 Edge based curvature and bending energy estimation . . . . . . . . . . . 101 A.4.2 Vertex based curvature and bending energy estimation . . . . . . . . . . 103 A.4.3 Comparing curvature estimation algorithms . . . . . . . . . . . . . . . . 104 A.5 Derivation of triangle-bead volume exclusion interaction . . . . . . . . . . . . . 108 A.5.1 Evaluating the integral . . . . . . . . . . . . . . . . . . . . . . . . . . . 108 A.5.2 Geometric interpretation of Eq. A.17 . . . . . . . . . . . . . . . . . . . . 111 A.5.3 Resolving numerical issues for coplanar cases . . . . . . . . . . . . . . . 115 A.6 Performance of bundle protrusion simulations . . . . . . . . . . . . . . . . . . . 119 Appendix B: Supporting information for Chapter 3 122 B.1 Estimating the relationship between mesh size and the maximum gradient of energy122 B.2 Membrane protein cluster nucleation . . . . . . . . . . . . . . . . . . . . . . . . 125 B.3 Membrane bending energy with protein inclusion . . . . . . . . . . . . . . . . . 127 Appendix C: Supporting information for Chapter 4 129 v C.1 Euler buckling study . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 129 C.2 Description of test cases . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131 C.3 Strain profiles . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 132 C.4 Parameterization . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133 C.5 Q(ŝ) and r(ŝ) in the geodesic models . . . . . . . . . . . . . . . . . . . . . . . . 136 C.6 Energies in the geodesic models . . . . . . . . . . . . . . . . . . . . . . . . . . 138 C.7 List of symbols . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 143 Bibliography 145 vi List of Tables 3.1 Membrane diffusion barrier simulation parameters. . . . . . . . . . . . . . . . . 52 3.2 Protein LLPS simulation parameters. . . . . . . . . . . . . . . . . . . . . . . . . 54 3.3 Curvature-induced protein sorting simulation parameters. . . . . . . . . . . . . . 57 4.1 The difference metrics Cr and Cd between each model and the dynamical model. . 80 4.2 The equilibrated filament energies E, external energies from bound linkers Eext, and total energy Etot for the three test cases. . . . . . . . . . . . . . . . . . . . . 80 A.1 The parameters used in simulation. . . . . . . . . . . . . . . . . . . . . . . . . . 96 A.2 Parameters for vesicle bundle deformation simulation. . . . . . . . . . . . . . . . 97 A.3 Parameters for filament bundle polymerization with different contact angles with the membrane . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 98 A.4 Time consumption of 0.2 s simulation time window when N = 7. Simulating 16.1 s in total took 1 d 6 h 22 min. . . . . . . . . . . . . . . . . . . . . . . . . . . 120 A.5 Time consumption of 0.2 s simulation time window when N = 10. Simulating 10.4 s in total took 3 d. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 120 A.6 Time consumption of 0.2 s simulation time window when N = 14. Simulating 4.01 s in total took 3 d. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 121 C.1 The deviation of the fitted prefactor a from 1 is shown for the five variational models tested. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 131 vii List of Figures 1.1 Existing models that involve membrane curvature generation in various cellular processes. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4 1.2 Schematic representation of the cell with various cytoskeleton structures. . . . . . 6 1.3 Membrane regulates actin network dynamics. . . . . . . . . . . . . . . . . . . . 8 2.1 Membrane manifold representation and deformation modes. . . . . . . . . . . . 17 2.2 An illustration of membrane as a chemical regulator in 2D space. . . . . . . . . . 28 2.3 Evolution of the vesicle shape is shown under increasingly hyperosmotic conditions. 31 2.4 Vesicle deformation by actin filaments crucially depends on the spatial organiza- tion of actin filaments. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 32 2.5 Filament bundle deforming the membrane. . . . . . . . . . . . . . . . . . . . . . 34 2.6 Membrane protrusion behaviors under various membrane rigidities. . . . . . . . 36 3.1 Geometric elements used in calculating a jk. . . . . . . . . . . . . . . . . . . . . 45 3.2 Membrane species diffusion with a barrier. . . . . . . . . . . . . . . . . . . . . . 52 3.3 Membrane protein phase separations. . . . . . . . . . . . . . . . . . . . . . . . . 55 3.4 Surface protein concentration sorting for membrane vesicles with various radii. . 58 4.1 The quantities used to specify a thin rod’s configuration in the Cosserat theory. . . 65 4.2 The cross-section of a filament with a bound cross-linker. . . . . . . . . . . . . . 76 4.3 A comparison of the variational and dynamical models. . . . . . . . . . . . . . . 81 4.4 The structural details of the chiral bundle simulation. . . . . . . . . . . . . . . . 83 4.5 Results from the chiral bundle simulation. . . . . . . . . . . . . . . . . . . . . . 86 4.6 A visualization of a contractile actomyosin network from a MEDYAN simulation. 89 A.1 Examples of fans around a vertex. . . . . . . . . . . . . . . . . . . . . . . . . . 99 A.2 The remeshing algorithm generates anisotropic mesh suitable for simulation. . . . 101 A.3 Ways of estimating the area at a vertex. . . . . . . . . . . . . . . . . . . . . . . . 105 A.4 Result of curvature estimate benchmark on a red blood cell. . . . . . . . . . . . . 107 A.5 The area element on a triangle facet. . . . . . . . . . . . . . . . . . . . . . . . . 109 B.1 Two equilateral triangles under a uniform energy gradient. . . . . . . . . . . . . 123 B.2 Large membrane mesh size reduces the accuracy. . . . . . . . . . . . . . . . . . 124 B.3 Snapshot of membrane protein LLPS with too large mesh size. . . . . . . . . . . 125 B.4 The area concentration profile among membrane cells with the phase-separated initial state. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 126 viii C.1 The Euler critical buckling force Fc for various filament lengths L̂ for five varia- tional models. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 130 C.2 A comparison of strain profiles along an equilibrated filament in different models. 134 ix List of Abbreviations ALPS Amphipathic lipid packing sensor Arp2/3 Actin-related protein 2/3 complex ATP Adenosine triphosphate BCR B cell receptor GC Geodesic Cosserat GEK Geodesic extensible Kirchoff GTP Guanosine-5’-triphosphate MEDYAN Mechanochemical dynamics of active networks MRR Membrane-related reactions NMII Non-muscle myosin II PLC Phosphorylated phospholipase C VASP Vasodilator-stimulated phosphoprotein WASP Wiskott-Aldrich syndrome protein WAVE WASP-family verprolin-homologous protein x Chapter 1: Introduction The cell is the fundamental unit of life, and its shape is defined by the plasma membrane. The membrane guards the frontier of the cell, allowing the cell to communicate with its sur- rounding environment [1]. Inside the cell there is a ubiquitous polymeric network known as the cytoskeleton. Crosslinked filament networks serve as a scaffold stiffening the cell [2]. With chemical energy supplies, the polymeric networks actively self-reorganize into various structures, converting chemical energy to mechanical work [3,4]. Through active remodeling, the cytoskele- ton can exert forces on the membrane, initiating the cell’s mechanical response. On the other hand, mechanical or chemical cues transmitted from the membrane can affect this cytoskeleton reorganization processes, complementing the cell’s mechanosensing process [5, 6]. Understand- ing how cytoskeleton works with the membrane to help the cell adapt to external environment is crucial in understanding how cell works. Computer simulations provide a powerful way to understanding such complex systems [7]. Knowledge of individual components in the system can be obtained from past theoretical, computational and experimental studies, and is used as the inputs of the simulation programs. The simulation programs then evolve the system based on fundamental physics and chemical principles, reproducing mesoscopic emergent behaviors that are difficult to obtain from analytical studies. Moreover, intermediate information that is generally inaccessible in experiments could 1 be easily extracted from computer simulations, providing more insights into the complex systems [4,8]. Over last two decades, efforts on simulation software on both the cytoskeleton systems and the membranes greatly enhanced understanding of these systems. With past advances, we aim to build a computation model that captures the essential mechanochemical properties of the entire cell, allowing cell dynamics predictions. In this thesis, we bring new simulation capabilities to the active matter simulation platform MEDYAN (Mechanochemical Dynamics of Active Networks) [9], including a new cytoskeleton- interacting membrane model with membrane protein dynamics. We also introduce a more ac- curate Cosserat-rod-based filament model, and use it to investigate symmetry breaking in actin bundles. In the remainder of the introduction, we first discuss the cytoskeleton and the membrane systems in greater detail, including their biological functions and existing models to describe them. Next, we discuss the cytoskeleton-membrane interactions and the efforts to simulate these processes. Finally, we will briefly outline the subsequent chapters in the thesis. 1.1 Biological membrane The membrane defines the boundary of a cell as well as many important organelles such as the mitochondria and the nucleus [10]. The membrane consists of a thin lipid bilayer as the main structural component as well as proteins attaching on the bilayer [11]. Mechanically, the membrane resembles a thin fluid film, resisting in-plane and out-of-plane deformations, which, in turn, influences its cellular functions. For example, membrane tension has been shown to influence exocytosis, endocytosis and cell motility [12, 13]. The membrane proteins can diffuse on the bilayer due to its in-plane fluidity [14]. The membrane proteins can be categorized into 2 a variety of classes, carrying out important cellular functions. For example, ion channels are transmembrane proteins that control flows of specific ions across the membrane and, in turn, control cell volume, action potential, exocytosis and many others [15]. Membrane proteins can also undergo clustering via 2D liquid-liquid phase separation (LLPS), a thermodynamic effect arising from intermolecular attraction [16–19]. Computation models for biological membranes spread across multiple orders of magnitude in terms of length scale and resolution, ranging from all-atom simulations to micrometer-scale meshless or manifold representations [20]. Membrane curvature generation processes have profound impact on cellular processes such as membrane trafficking and endocytosis [21, 22]. Many different membrane proteins influ- ence membrane curvature with various pathways [23]. For example, the BIN-Amphiphysin- Rvs (BAR) domain proteins locally curve the membrane by both adapting the lipid bilayer to its intrinsic curvature [24], and by inserting hydrophobic helices into the lipid layer [23, 25]. Conversely, many curvature generating proteins are known to sense the membrane curvature as well. In the BAR domain protein example, the helix insertion was also identified to be the key factor for curvature sensing processes [25]. Advances in computer modeling have enabled bet- ter understanding of mechanisms underlying membrane curvature sensing and generation (see Fig. 1.1) [23, 26–29]. However, how the curvature sensing and generating effects synergistically reinforce each other and modulate the membrane shape is still not well understood [23]. 1.2 Cytoskeleton Cytoskeleton exists ubiquitously in eukaryotic cells, consisting of interconnected poly- mer networks with regulatory proteins [1]. The cytoskeletal networks can actively undergo non- 3 Figure 1.1: Existing models that involve membrane curvature generation in various cellular pro- cesses. (A) Membrane budding process during endocytosis [26, 30]. (B) Tubular membrane structure formation and stabilization [28, 31]. (C) Nuclear envelope topology change [32]. (D) Membrane deformation by caveolae [33, 34]. (E) Membrane protrusion by actin force [27, 35]. (F) Mitochondrial fission [36, 37]. Image reprinted from Ref. [23] from Molecular Diversity Preservation International under Creative Common Attribution License. 4 equilibrium reorganization, powered by chemical energies [38]. For example, with ATP supplies, an actin filament — a major type of polymer in the cytoskeleton — can have net growth on the barbed and and net destruction on the pointed end at steady state [39,40]. The energy-consuming polymerization of the filaments can also be converted to useful work under external loads [41]. Another example is the non-muscle myosin II (NMII), which a crosslinking protein with heads that can bind to actin filaments. NMII heads can consume ATP to move along actin filaments [42], which may create stress in cytoskeleton and lead to network contraction [9, 43]. With highly dy- namic components, the cytoskeletal networks can self-organize into various structures (Fig. 1.2), and control important cellular activities such as cell stiffening, migration, transport and mechan- otransduction [38, 44–48]. Computer simulation packages have been developed for mechanochemical models of cy- toskeletal networks. Examples include MEDYAN [9], Cytosim [50], AFiNeS (Active Filament Network Simulation) [51], etc. In this work, we base our work on the MEDYAN platform. In MEDYAN, diffusing chemical species in each compartment are considered “well-mixed”, but the locations for force bearing species (such as actin filaments and bound myosin motors) are explicitly modeled and tracked [9]. MEDYAN simulations are based on iteratively interleaving stochastic draws of chemical events, using the Gillespie algorithm, followed by a mechanical equilibration step. Importantly, chemistry influences mechanics and vice versa, creating intricate mechanochemical feedbacks. In MEDYAN, polymeric filaments are modeled as straight, compressible cylinders con- nected at hinge points [9]. Since these filaments have high aspect ratio (an actin filament can reach several micrometers in length and has only ∼ 7nm in diameter [52]), this simplified ap- proach captures salient filament properties such as stretching and bending rigidities, and is com- 5 Figure 1.2: Schematic representation of the cell with various cytoskeleton structures. (i) The cell cortex. (ii) A contractile stress fiber. (iii) The lamellipodium. (iv) The filopodium. Image reprinted from Ref. [49] with permission from the American Physiological Society. 6 putationally cheap enough to allow for micron-scale actomyosin simulations to reach a timescale of ∼1 hour. However, this model neglects finite width of the filaments and omits the chiral asymmetry of helical actin filaments [53], so simulations using it cannot reproduce nor explain emergent chiral symmetry broken behavior observed in experiments [54]. To address this is- sue, in this work, we present new filament models in MEDYAN based on Cosserat rod theory, capturing all allowed deformation modes, while minimizing the impact on computation times. 1.3 Cytoskeleton-membrane interactions A complete cellular mechanochemical sensing process often involves the membrane, a mostly inactive flexible surface, cooperating with the cytoskeleton, a highly active dynamic poly- mer network [55–57]. Membranes can interact with the cytoskeleton mechanically via repulsion and adhesion [58]. Cells with the plasma membrane attached to a dense cross-linked actin cor- tex have their boundaries significantly rigidified [12]. On the contrary, disruption of filament- membrane attachments leads to cellular blebbing [59,60]. Membrane proteins also participate in numerous signaling processes, regulating, in particular, many cytoskeletal interactions (Fig. 1.3). For instance, the membrane attached WASP-VASP complex regulates Arp2/3 activation, which is an actin branching protein essential in lamellipodia formation [61]. From a different perspective, membrane receptor proteins and their spatial organizations play an indispensable role in cellular signaling processes, which modulate, for example, the formation of immunological synapse [63, 64]. Clustering of receptors may alter the signaling cascade and affect cytoskeleton dynamics. For example, transmembrane complex nephrin-NCK- N-WASP clustering regulates actin assembly [65, 66]. Reciprocally, the cytoskeleton dynamics 7 Figure 1.3: Membrane regulates actin network dynamics. (A) Membrane lipids affect profilin activity. Profilin preferably binds to PI(4,5)P2, allowing for more free actin monomers locally, while phosphorylated phospholipase C (PLC) hydrolyzes PI(4,5)P2 and allows for more profilin- bound actin monomers locally. Free actin monomers preferably nucleate with Arp2/3 complex to form branched networks, while profilin-bound actin monomers prefer formin-mediated fila- ment polymerizations. (B) Small activated GTPases of the Rho superfamily proteins attach on the membranes, activating WASP/WAVE factors, which modulates Arp2/3 activity. F-BAR pro- teins interact with WASP, regulating actin polymerization. These interactions lead to processes described in the text boxes. (C) Small activate GTPases of the Rho superfamily on the membrane can bind to formins, activating actin polymerization. F-BAR proteins also regulate formin activ- ities, which, in turn, modulates actin activities and drives membrane protrusion and cytokinesis. Formins binding directly to the membrane can be found in eukaryotes that lack formins with obvious Rho-binding domains, such as plants. Image reprinted from Ref. [62] with permission from Rockefeller University Press. 8 in membrane’s vicinity can influence the distribution of membrane proteins as well, facilitating or hindering their clustering. For example, the spatial distribution of the transmembrane recep- tor CD44 is guided by the underlying cytoskeleton dynamics [67]. Moreover, these complex interactions may further deform the membrane into highly curved shapes, and, in turn, affect the dynamics of curvature-sensing proteins. Specifically, How membrane-actomyosin dynamics, driven by far-from-equilibrium processes, and the passive intermolecular interactions synergisti- cally regulate the aggregation and disassembly of surface receptors remains a central question at the interface of physical chemistry and cell biology. A simulation framework that captures these complex surface-receptor/membrane/cytoskeleton interactions may greatly improve the under- standing of mechanism of surface protein phase behavior. Efforts of using computational models to study cell periphery dynamics involving membrane- cytoskeleton interactions have been made in the past decades. For example, both analytical and computational methods have been used to study filopodia and lamellipodia systems [68–70]. A recent work combined branched actin network simulation with hemisphere-shaped membrane endocytic pit, and emphasized cytoskeleton’s role in initiating endocytosis [71]. However, it is still desirable to develop a detailed, flexible, 3D membrane model which interacts with cytoskele- tons. In this thesis, we will introduce our new membrane model in MEDYAN, capturing salient properties such as membrane surface proteins and membrane-filament interactions. This work enables cytoskeleton mechanochemical simulations enclosed in a movable membrane domain, paving way for whole cell modeling. 9 1.4 Outline of Thesis In Chapter 2, we present a new computational model for simulating membranes and their interactions with the cytoskeleton. We introduce a mesh-based membrane model in MEDYAN, capturing the membrane’s elastic properties, membrane-filament steric interactions and volumet- ric effects such as the osmotic pressure. The membrane-MEDYAN simulations reveal how the growing actin bundles generate tubular membrane protrusions. We discuss our investigations on the protrusion initiation and dynamics under various filament bundle geometries, membrane rigidities and local G-Actin concentrations. Using simulations, we found that the bundles’ protru- sion propensities depend on the synergy between bundle thickness and inclination angle at which the bundle approaches the membrane. The new membrane model paves the way for whole-cell simulations using MEDYAN. In Chapter 3, we present a novel way to simulate surface protein dynamics using a compartment- based reaction-diffusion scheme on membrane meshes. We show how reaction rates and diffusion coefficients for proteins in a potential field can be translated to reaction/hopping propensities on unstructured 2D meshes. Then, we show membrane proteins undergoing thermodynamics driven LLPS, which can be studied using this model by incorporating the protein inter-molecular in- teractions. Moreover, utilizing the curvature-dependent membrane binding energy in the protein potential field, we demonstrate curvature-induced protein curvature sorting on membrane vesi- cles with various sizes and analyze the sorting strength under various membrane protein elastic properties. In Chapter 4, we introduce new variational models for cytoskeletal filaments. In order to simulate large filament networks that incorporate stretching, shearing and twisting of filaments 10 with finite radius, we propose new filament models based on the Cosserat rod theory. We compare the cylinder-bead filament model in the original MEDYAN, a polynomial spline-based Cosserat filament model, as well as a Cosserat model where geodesic curves in the configurational space is used, and find that the last model both captures all the relevant degrees of freedom and allows for arbitrarily large deformations without systematic error. We implement new procedures necessary for these models to work properly with MEDYAN, and show that the new models reveal chiral symmetry breaking for filament bundles with torsional compliance. In chapter 5, we provide the summary of the work in this thesis and discuss future directions of research. 11 Chapter 2: Simulating deformable vesicles containing complex cytoskeletal net- works This chapter and Appendix A are adapted with permission from: Haoran Ni and Garegin A. Papoian (2021) Membrane-MEDYAN: simulating deformable vesicles containing complex cy- toskeletal networks. The Journal of Physical Chemistry B, 125(38), 10710-10719. Copyright 2021 American Chemical Society. 2.1 Introduction Cytoskeletal fibers, along with regulatory molecules, such as α-actinin cross-linking nearby actin filaments, and myosin walking on the actin filaments, work together to establish cell’s main mechanochemical engine, continuously consuming energy. The plasma membrane, on the other hand, is mostly inactive. However, it defines the boundary of a cell, playing an important role in being a mechanical and chemical window to the outside world, regulating cytoskeletal functions [72]. Mechanically, the membrane is a thin fluid film, resisting in-plane and out-of-plane defor- mations, which, in turn, influences its cellular functions. For example, membrane tension has been shown to influence exocytosis, endocytosis and cell motility [12, 13]. Membranes can in- teract with the cytoskeleton mechanically via repulsion and adhesion [58]. Cells with the plasma 12 membrane attached to a dense cross-linked actin cortex have their boundaries significantly rigid- ified [12]. On the contrary, disruption of filament-membrane attachments leads to cellular bleb- bing [59, 60]. Membrane proteins also participate in numerous signaling processes, regulating, in particular, many cytoskeletal interactions. For instance, the membrane attached WASp-VASP complex regulates Arp2/3 activation, which is an actin branching protein essential in lamellipodia formation [61]. Computer simulations can shed light on mechanisms of membrane-cytoskeleton interac- tions, building on and complementing extensive experimental observations of these processes [72]. A variety of such computational models have been proposed during the last two decades. Both analytical and computational models were used to study the filopodia formation and prop- erties [68, 69]. To study lamellipodial protrusions, the membrane leading edge was modeled as a quasi 1D curve, interacting mechanochemically with a branched actin filament network [70]. Following up on these advances, it is desirable to develop a flexible, scalable, 3D model for mem- branes and their interactions with cytoskeleton that has detailed mechanochemical components essential for simulating whole-cell dynamics. With this goal in mind, we based our work on MEDYAN (Mechanochemical Dynamics of Active Networks, a software designed for carrying out detailed cytoskeletal simulations [9]), extending it such that cytoskeletal networks can be enclosed within a movable membrane domain. MEDYAN can simulate cytoskeletal behaviors at high spatial and temporal resolutions. It couples cytoskeletal network mechanics with spatially resolved chemical dynamics, enabling studies of cytoskeletal processes at the micrometer scale and the time scale of tens of minutes [9]. Simulations using MEDYAN have shed light on the turnover and treadmilling dynamics for actin filaments [73], the stability and shape transformations of actin bundles [74] and the entropy pro- 13 duction and cytoskeletal avalanches (cytoquakes) of evolving actomyosin networks that are far from equilibrium [4,75,76]. Currently, however, the enclosing domain for cytoskeletal evolution in MEDYAN is rigid and immovable, imposing inflexible steric restriction on cytoskeletal fila- ments. In this work, we incorporate into MEDYAN an explicit membrane model, which accounts for membrane’s salient mechanical properties, volumetric effects such as osmotic pressure, as well as filament-membrane interactions. The new membrane model paves the way for whole-cell simulations using MEDYAN. In this work, we applied membrane-MEDYAN to simulate cytoskeletal networks enclosed in a vesicle, finding that actin network’s architecture plays an important role in generating mem- brane tubular protrusions. Furthermore, our analysis indicates that bundle protrusion initiation sensitively depends on local G-Actin concentration, bundle thickness, membrane rigidity and the inclination angle of the bundle approaching the membrane. The remainder of the chapter has the following organization. In Section 2.2, we introduce the new membrane model, including its discrete representation, the mechanical energy model and the membrane-cytoskeleton interactions. In Section 2.3, we discuss the application of membrane- MEDYAN to several relatively simple model systems. In particular, we report on our simulations of vesicles containing various actomyosin networks, exploring conditions that lead to formation of successful protrusions. Finally, in Section 2.4, we summarize our work, and further elaborate on the limitations and potential applications of the new model. 14 2.2 Methods 2.2.1 MEDYAN Simulations The new membrane model was incorporated into MEDYAN. The latter can be used to simulate active matter systems using a compartment-based reaction-diffusion scheme, where dif- fusing chemical species are considered “well-mixed” in each compartment, but the locations for force bearing species (such as actin filaments and bound myosin motors) are explicitly modeled and tracked [9]. MEDYAN simulations are based on iteratively interleaving stochastic draws of chemical events, using the Gillespie algorithm, followed by a mechanical equilibration step. Importantly, chemistry influences mechanics and vice versa, creating intricate mechanochemical feedbacks. 2.2.2 Membrane mechanics and discretization Typically, the phospholipid bilayer has a thickness of 3− 7 nm [77]. At the scale of a whole cell (1− 10 ţm), the thickness of the plasma membrane is much smaller than its lateral dimensions. Therefore, we model the plasma membrane as a 2D Riemannian manifold, M , embedded in the 3D Euclidean space. Some of the mechanical properties of the lipid bilayer, arising from in-plane and out-of-plane deformations and from interactions with the cytoskeleton, are mapped onto this 2D manifold. We note here that from the perspective of a point on the surface, we use the term “lateral directions” to denote the directions in its tangent plane. Next, we discretize the 2D surface M into a triangulated meshwork M̂ (Fig. 2.1a). The mesh is comprised from sets of vertices V , edges E and triangles T . Each vertex vi ∈ V is a 15 representative point on the original surface. The 3D coordinates of all vertices control the shape of the surface. Each triangle t ∈ T is a set of 3 vertices {vi,v j,vk} ⊂ V . These piecewise linear triangles tile the entire surface of M̂ , which, in turn, approximates the original manifold M . Each edge e is a set of 2 vertices {vi,v j} ⊂ V , and the set of all edges E := {e = {vi,v j} : e ⊂ t for some t ∈ T}. We define the 1-ring neighbor vertices Nv,1(v) := {vi ∈ V : {v,vi} ∈ E}, and the 1-ring neighbor triangles Nt,1(v) := {t ∈ T : v ∈ t}. Several other constraints are added to the structure of the mesh to enforce a manifold-like structure. See the SI for more information. Below, all membrane related potentials take a discretized form as well, using coordinates of the vertices as corresponding arguments for the potentials and force functions elaborated below. Excluding interactions with the cytoskeleton, we define the total mechanical energy for the membrane as the sum of the following potentials (see Fig. 2.1b top-left, top-right, bottom-left), Emembrane = Earea +Etension +Ebending +Evolume, (2.1) where Earea is the area elastic energy, Etension is the surface tension energy, Ebending is the bending energy, and Evolume is the free energy associated with the 3D volume enclosed by the membrane, which can arise, for example, as a result of the intracellular osmotic pressure. The discretized form of these free energies are functions of mesh vertex coordinates, hence, one can compute the forces on vertices, with fi j = −∂E/∂x j(vi), where fi j indicates the j-th component of the force on vertex vi, x j(vi) is the j-th coordinate of vertex vi, for vi ∈V and j ∈ {1,2,3}. In the above outlined discretization scheme, mesh vertices do not have to represent the material points: the time evolution of an individual vertex does not represent the trajectory of any lipid molecule in the membrane. Therefore, one may also freely remesh the original surface at any 16 Figure 2.1: (a) The plasma membrane is represented as a 2D manifold M , and is implemented as a triangulated mesh M̂ . (b) The potential terms in our membrane model are illustrated. The shapes with a dashed stroke indicate allowed deformations that lead to: (top-left) surface tension; (top-right) membrane bending rigidity; (bottom-left) volumetric potentials in case of closed sur- face; (bottom-right) membrane-filament repulsion. The actin filaments are shown in red. 17 time as long as the resulting shape is a good approximation of the original shape. Alternatively, we can enforce the vertices to be material points, where each vertex represents a patch of lipid molecules on the membrane that always stay together. In this case, the local area elasticity can be incorporated naturally in the membrane model, which will be discussed below. Our membrane model supports both options. In addition to closed surfaces, we can also represent membranes with fixed boundaries. The boundaries can optionally connect the membrane surface to an implicit lipid reservoir, allowing the total number of lipid molecules to change. Membrane area elasticity A piece of free open membrane spontaneously relaxes to an equilib- rium area that has zero surface tension [78]. Assuming homogeneity of lipid composition of the membrane, in the continuum limit, the membrane is considered to have a constant equilibrium mass density (mass per unit surface area) ρ0, so the free energy density (free energy per unit surface area) earea can be written in a quadratic form earea = karea 2 ( ρ ρ0 −1 )2 , (2.2) where ρ is the local density on the surface, and karea is a positive constant with dimension MT−2, where M and T are the dimensions for mass and time, respectively. Total elastic energy is then, Earea = ∫ M eareadS. For a membrane not connected to a lipid reservoir, the mass conservation equation ∫ M ρdS = m must be satisfied, where m is the constant total mass. In these integrals, dS is the surface area element. Assuming that vertices represent material points, there is a natural discretization of surface 18 elastic energy using Eq. 2.2. Each vertex vi carries a fixed amount of lipid molecules around it, and, therefore, has a well-defined equilibrium area A0(vi) around it. The actual area around a vertex A(vi) is defined as a third of the sum of areas of its 1-ring neighbor triangles, so A(vi) = ∑t∈Nt,1(vi)A(t)/3, where A(t) is the area of triangle t. Assuming that the surface area density in the vicinity of a vertex is uniform, the area elastic energy of the discretized membrane can be written as Êarea = ∑ v∈V karea 2A0(v) (A(v)−A0(v)) 2 , (2.3) where the “hat” on E indicates that it is a discretized form. On the other hand, if the vertices only parametrically represent the surface, not being con- nected to material points, the area elasticity cannot be naturally discretized around each vertex. However, if the membrane is not buffered by a lipid reservoir, we can use the mass conservation equation, with the assumption that the surface density is uniform across the whole membrane, to obtain the following relation, Êarea = karea 2A0(M̂ ) ( A(M̂ )−A0(M̂ ) )2 , (2.4) where A(M̂ ) = ∑v∈V A(v) is the area of the discretized membrane, and A0(M̂ ) is the equilibrium area of the membrane. One can check, using the Cauchy-Schwarz inequality, that the energy in Eq. 2.4 is a minimum of the energy reachable in Eq. 2.3, if A0(M̂ ) = ∑v∈V A0(v). The uniform density assumption is valid, because at the smallest time resolution that we are interested in (∼1 ms), the inhomogeneity of density on the membrane is quickly relaxed throughout the membrane due to its in-plane fluidity [12, 79]. 19 In fact, the value of karea for biomembranes is so large, that often the membrane area in- compressibility is assumed, and the Lagrange multipliers are used in calculations to fix the total area [80]. However, since we employ energy minimization as the main mechanical simulation method, we use Eq. 2.3 or Eq. 2.4 to enforce a softer constraint, by allowing the area of mem- brane to change but with large energy penalty, instead of strictly fixing the area. Membrane surface tension Many types of cells can actively maintain the membrane tension within a certain range via various biological processes, such as membrane undulation and caveo- lae formation and dissolution, which function as implicit lipid reservoirs [12]. The experimentally measured tensions typically range from 0.01 mNm−1 to 0.04 mNm−1 [81]. If an implicit lipid reservoir is considered, the total mass of the membrane is no longer conserved, and in this case, the vertices in our model will not be associated with material points. Assuming a constant surface mass density and a constant chemical potential of the lipid reservoir, the total area-dependent energy is reduced to a linear form, Etension = γA(M ), (2.5) being further discretized to Êtension = γA(M̂ ) where γ is the surface tension, a property of the external reservoir. Membrane bending elasticity The bending energy can be described using the Helfrich Hamil- tonian [82, 83], Ebending = ∫ M dS [ 2κH(H −H0) 2 +κGK ] , (2.6) 20 where κH and κG are the bending modulus and the saddle-splay modulus respectively, H is the mean curvature, H0 is the spontaneous curvature and K is the Gaussian curvature. According to the Gauss-Bonnet theorem, the contribution from the second term (the Gaus- sian curvature term) is constant with fixed membrane topology and boundary conditions, so we can ignore the second term for most purposes. We discretize the first term in Eq. 2.6 on a per- vertex basis. Êbending = 2κH ∑ v∈V (Ĥ(v)−H0(v))2A(v), (2.7) where Ĥ(v) and H0(v) are the mean curvature and the spontaneous mean curvature estimated around vertex v, and are assumed to be uniform in the neighborhood of the vertex with area A(v). Many different ways of estimating the surface curvature on a triangulated surface mesh are discussed in [84, 85]. Based on our numerical experiments, we decided to allow for two different ways of estimating the mean local curvature at a vertex in MEDYAN, which provide good numerical stability for the systems that we studied. The first way is described in the Surface Evolver [86, 87], Ĥ(vi) = 〈∇iA,∇iV 〉 2〈∇iV,∇iV 〉 , (2.8) where A is the total area of the membrane, V is the total volume enclosed by the membrane, ∇i means “taking the gradient with respect to coordinate of vertex vi”, and 〈·, ·〉 denotes taking the inner product of the vectors. This definition gives a continuous and signed curvature value, the sign depending on the orientation of the surface, because the volume calculation is orientation- dependent. Therefore, this curvature estimate is suitable for bending energy computation when the spontaneous curvature H0 is not zero. In the continuum limit, ∇A and ∇V are both parallel to the normal direction, but computed numerically on a discretized mesh. Hence, there can be 21 misalignments between these vectors. Under extreme conditions these two vectors can even become almost perpendicular, leading to unphysical geometries. When H0 = 0, the sign of the curvature no longer matters in Eq. 2.7, and we rely on a way to directly estimate the unsigned curvature as, Ĥ(vi) = ‖∇iA‖ 2‖∇iV‖ . (2.9) This equation provides better numerical stability under extreme conditions. For detailed discus- sion on how the bending energy was implemented, see the SI. The volume for a closed discretized surface mesh can be computed as V (M̂ ) = ∑t∈T V (t), where V (t) is the signed volume of the tetrahedron formed by the triangle t and an arbitrary fixed point p ∈ R3, and the sign depends on the orientation. It is often convenient to use the origin of the coordinate system as the fixed point. In-plane shearing of the membrane is ignored We do not explicitly include the in-plane shearing stress of the membrane in our model. Because the membrane behaves as a fluid in the lateral directions, the in-plane shearing stress cannot be sustained, while there can be a dynamic stress, which depends on fluid’s properties. In MEDYAN, mechanical energy minimization steps imply that the system is nearly equilibrated [9]. Therefore, we also ignore the dynamic shearing stress. Volumetric energy A membrane vesicle is also resistant to a change of its enclosing volume. The following quadratic expression captures the volume dependent free energy near the equilib- rium volume, Evolume = kvolume 2V0(M ) (V (M )−V0(M ))2 , (2.10) 22 where V (M ) and V0(M ) are the volume and the equilibrium volume enclosed by the membrane M respectively, and kvolume is a positive constant, with dimension ML−1T−2, where M, L and T are the dimensions for mass, length and time, respectively. Next, we show how this formula can be derived using the osmotic pressure difference across the membrane vesicle. Assuming a constant osmotic pressure outside the vesicle, the free energy due to osmotic pressure inside an enclosed volume is, Eosmotic =− ∫ V V ∗ ∆Π(V )dV =− ∫ V V ∗ (Πin(V )−Πout)dV, (2.11) where Πin and Πout are the osmotic pressure inside and outside the vesicle respectively, ∆Π = Πin −Πout, V is the volume enclosed by the membrane and V ∗ is an arbitrary reference volume. Under an assumption that the membrane is perfectly permeable to water and impermeable to solutes, the minimum osmotic energy occurs when V =V0 and Πin(V0) = Πout. Under further assumptions that the total mole fraction of the solutes is far less than the mole fraction of the solvent, and that the solutes are well mixed in the system, one can write Πin = NkBT/V , where N is the total number of solute molecules in the vesicle. In this case, V0 = NkBT/Πout, and by choosing V ∗ =V0, the energy can be evaluated as Eosmotic = Πout [ −V0 ln V V0 +(V −V0) ] ≈ Πout 2V0 (V −V0) 2, (2.12) where the approximation is the result of expanding the energy to the second order with respect to V around V0. 23 Additional discussion of the membrane force field Our membrane model is based on Helfrich’s scheme, which assumes that the lipid molecules are perpendicular to the membrane surface [82]. However, lipid tilting may also contribute to the total free energy, which was shown to be im- portant at length scales of lipid molecules themselves [88]. In our current model, we are not including the tilting effect because the resolution of the membrane mesh in our model (i.e., the triangle element length) is typically 20− 70 nm, which is much larger than the length scale at which tilting plays an important role. Our current model also does not include a repulsive potential among the membrane ele- ments. Although this may hypothetically allow unphysical geometries for folded membranes, in our simulations reported below, we have not observed any membrane self-crossing. In part this is explained by the steric interactions between the membrane and actomyosin networks that engen- der a positive tension, preventing crossing of membrane’s segments. In future, a self-avoidance potential can be naturally incorporated if self-crossing becomes a problem, however, at the ex- pense of added computational effort. For 2D manifold, all free energy terms mentioned above, except for the neglected in-plane shearing energy, are parametrization invariant, which means that the energy of the surface de- pends only on the shape of the surface. How the surface is triangulated or how the surface coordinates are chosen only affect the discretization error. More specifically, in the absence of membrane shearing energy, as long as the triangulations represent the same shape of the surface up to some precision (the similarity between two surfaces can be measured using the Hausdorff distance [89]), the total discretized free energies computed on those triangulations should be approximately the same. This is, in fact, very useful in a simulation, because it allows us to re-triangulate the membrane meshwork to improve the triangulation quality, without changing 24 membrane’s physical behaviors. This approach effectively addresses the problem of the quality of triangulation becoming poor over time due to membrane’s large deformations, which, in turn, leads to undesirable numerical instabilities. 2.2.3 Adaptive surface remeshing The aforementioned membrane potentials (especially bending energy, that is based on a discretized estimate for curvature) suffer numerical issues when the membrane deformation be- comes large. To resolve this problem, we rely on the method of triangulated surface remeshing, which essentially optimizes the membrane quality without changing its shape [90, 91]. The quality of the mesh is related to the following: the triangle shape qualities and the dihe- dral qualities. For best numerical stability, the triangles used in the mesh should be as equilateral as possible, while the dihedral angles between the triangles sharing an edge should be as close to π as possible (or the angle between oriented normal vectors of the two triangles are as close to 0 as possible). Our implementation of such an algorithm was inspired by [90, 91]. We apply local opera- tions on the meshwork in an iterative manner to optimize both triangle and dihedral quality. With this algorithm, we are able to get an isotropic surface triangular mesh with good quality, where the vertex density adapts to the local curvature, making numerical simulations accurate for high curvature regions, while being efficient for low curvature regions. See the figures in the SI for an example of the remeshing procedure. In MEDYAN, the adaptive mesh quality optimization for a membrane is carried out after every energy minimization. We rely on the Marching Tetrahedra algorithm as the mesh gen- 25 erator [92], which is fast, robust and easy to implement. However, the quality of the mesh by Marching Tetrahedra may lead to triangles that are far from being equilateral. Therefore, the mesh optimization algorithm is essential for achieving numerical stability. 2.2.4 Interactions between the membrane and the cytoskeleton 2.2.4.1 Membrane-filament repulsion Without tethering proteins connecting the cytoskeletal filaments to the plasma membrane, the filaments and the membrane sterically repel each other (Fig. 2.1b bottom-right). A hypothet- ical repulsive interaction between filament beads and membrane vertices would be insufficient in this case, because our simulations showed that the filaments are able to occasionally pierce through the triangle. Therefore, we developed an excluded volume interaction via repulsion be- tween filament’s tip and a meshwork triangle, where essentially every point on the surface of the membrane interacts with the tip. The interaction energy between a given filament tip and membrane takes the following form, Eexcl = kexcl ∫ M dS 1 dn = kexcl ∑ t∈T ∫ t dS 1 dn , (2.13) where kexcl is the force constant to the interaction and n describes the steepness of the potential, and d is the distance between the filament tip and a point on the membrane. This form of the potential generalizes the potential for segmented interactions from [93] to higher dimensional discrete elements. We found that choosing n= 4 produces a repulsive force field that is steep enough while still amenable to analytical evaluation of forces. For detailed derivation of this volume exclusion energy, see the SI. 26 The force acting on the filament tip from this repulsive interaction is f =−∇Eexcl, where ∇ means “taking the gradient with respect to the coordinate of the filament tip”. The reaction force (repulsive forces on the membrane by the filament tip) is balanced by the load on the membrane when the mechanical energy is minimized. Therefore, we directly use the value of f to be the load force when a filament polymerizes against a membrane. The polymerization of a filament near the membrane ratchets the membrane thermal fluctuation [94], and statistically work is done against the membrane load, resulting in membrane protrusion or deformation [41]. In the absence of explicit modeling of membrane fluctuation, we use the value of the load force f to compute a scaled filament polymerization rate [9, 41], kpoly,eff = kpolye 〈 f ,δ 〉 kBT , (2.14) where 〈·, ·〉 indicates an inner product, and δ is a vector pointing in the direction of polymeriza- tion. The average monomer length projected along an actin filament is, ‖δ‖ ∼2.7 nm. This cou- pling introduces a non-linear feedback that allows instantaneously generated mechanical stresses to locally modulate chemical dynamics. A schematic diagram of membrane-filament interaction can be seen in Fig. 2.2. 2.2.4.2 Membrane and diffusing species Diffusing large molecules, such as G-actin monomers, cannot move across the membrane without assistance. Therefore, the membrane must confine the space where the protein diffu- sion and chemical reactions take place. In a compartment based reaction-diffusion system, the effect of membrane confinement of reaction volume is achieved by full or partial deactivation 27 Figure 2.2: An illustration of membrane as a chemical regulator in 2D space. The membrane disallows big diffusing molecules to move across, such as G-Actin molecules or the α-actinin linkers. Therefore, the reaction volume is defined by the region inside the membrane (approx- imated by the straight dotted lines cutting through the compartments). The polymerization of filaments against the membrane is, however, possible but with reduced rates due to the Brownian ratchet effect. The actual implementation is carried out with 3D geometrical elements. 28 of reaction networks in affected compartments. As shown in Fig. 2.2, a completely deactivated compartment does not allow chemical reactions or diffusion with neighboring compartments. A partially deactivated compartment, however, works as a normal compartment, but the reaction and diffusion rates are scaled by the diminished volume as well as the diminished contact area between neighboring compartments. To simplify computation, if a part of a membrane cuts through a certain compartment, we fit the shape of that part of the membrane into a plane, and perform a planar cut through that compartment to obtain the sliced volume and the sliced area on each side of the compartment. This latter method introduces a small inaccuracy in sliced volume computation. To estimate the corresponding error, we initialized a 0.8 ţm radius spherical vesicle at the center of a 6×6×6 compartment grid, where each compartment is a cube of size 500 nm. For each partially activated compartment sliced by the membrane mesh, we first estimated the actual sliced volume using a Monte-Carlo integration with 10000 samples. We found that the average relative error between the planar approximation of the sliced volume and the more accurate Monte Carlo estimate is below 4%. If needed, this error can be further reduced by choosing a smaller compartment size. 2.2.5 Computational efficiency Every simulation in this work was run on a single CPU core. Detailed benchmarks on the performance of perpendicular bundle protrusion simulations (discussed in the next section) are reported in the SI. These benchmarks indicate that mechanical force relaxation is the most time-consuming process in our simulations. Hence, parallelization of the mechanical energy computation will enable accessing significantly larger time and length scales. 29 2.2.6 Code availability The source code for this work is published with MEDYAN version 5 on http://medyan. org, where example input files for simulating filament-vesicle interactions are provided in the examples directory. Additionally, simulation input files and trajectories for results presented below have been archived in Digital Repository at the University of Maryland [95]. 2.3 Results and Discussion We first applied the new force field to investigate how a spherical vesicle would deform as its volume starts to shrink under hyperosmotic conditions. The latter causes a mismatch of internal and external pressures, which affects our new volumetric term, leading to corresponding reduction of system’s volume. We started simulations with an initially spherical vesicle having a radius of 1 ţm, applying potentials given in Eqs. 2.4, 2.7 and 2.12. The equilibrium area, A0, was fixed to be the initial spherical surface area. We continuously decreased the equilibrium volume V0 over time, which effectively creates an increasingly hyperosmotic chemical environment, since V0 = NkBT/Πout under dilute conditions. Several snapshots of the simulation trajectory are shown in Fig. 2.3, indicating that un- der a constant area constraint, the membrane is unable to maintain its original spherical shape, crumpling as the vesicle volume shrinks. The local surface deformations, however, are smooth, because the bending potential prevents formation of sharp features where the mean curvature would be too large. We also notice that, over time, the regions with high convex curvature (as shown in blue in Fig. 2.3) become more distinct and highly localized. Next we investigated how two classes of actomyosin architectures interact with the vesi- 30 http://medyan.org http://medyan.org Figure 2.3: Evolution of the vesicle shape is shown under increasingly hyperosmotic conditions. Scale bar: 1 ţm. The red-white-blue coloring scale marks the local mean curvature, with red being the most concave and blue being the most convex. The membrane surface becomes more crumpled over time, and the bending stress is balanced by the forces arising from volume and area potentials. Using adaptive remeshing, the mesh is sampled more densely where the absolute values of local curvatures are higher. cle membrane boundary. Our simulations of vesicles containing disordered actomyosin networks, which are not bundled, indicate that randomly initialized filaments growing toward the membrane tend to bend instead of making a membrane protrusion, continuously inducing small membrane ruffles (Fig. 2.4a). At a steady state, the filaments accumulate just beneath the membrane sur- face. On the other hand, as displayed in Fig. 2.4b, the initially bundled filaments grow toward two poles of the vesicle, forming strong tubular protrusions. Interestingly, in some cases, a few filaments grow away from the axis of the bundle, become eventually buckled, hence, resembling the behavior of filaments in disordered actomyosin networks. Finger-like membrane protrusions, such as the filopodia, are ubiquitous structures in eukaryotic cells [96, 97]. Although the bio- logical functions of the filopodia were widely explored [97–99], a clear understanding of how filopodia form as a result of membrane-cytoskeleton interactions is still lacking. Overall, these results demonstrate that distinct seeding patterns lead to different actomyosin network architec- tures, where formation of bundles is crucial for generating tubular membrane protrusions. To study what bundle properties most facilitate protrusion, we explored how the tubula- 31 Figure 2.4: Vesicle deformation by actin filaments crucially depends on the spatial organization of actin filaments. The color codes are the same for two images: Cyan: Membrane, Red: Actin filaments, Green: α-actinin cross linker, Blue: Myosin motor. Filaments are randomly initial- ized in (a), and initialized as a bundle in (b). The boxed images on the bottom right show the initial filament arrangements for each condition. Bundled filaments can generate tubular protru- sion, while filaments in disordered actomyosin networks accumulate under the membrane, being largely parallel to the surface. 32 tion process depends on the number of actins filaments and G-Actin concentrations. To sus- tain stable tubular growth, the number of filaments, N, should satisfy kpoly[A]exp(−‖ f‖‖δ‖ NkBT ) ≥ kdepoly [68, 69], where ‖ f‖ = 2π √ 2κγ , which is the load force at the tip of a stable filopodium. This imposes a minimum G-Actin concentration threshold which increases exponentially as the number of actin filaments decreases. For example, a bundle with 7 actin filaments requires at least 3.9× 10−4 mol/m3 of G-Actin at the tip to sustain protrusion against the membrane with γ = 0.02pNnm−1 and κ = 100pNnm, but with 1 filament alone, the concentration has to be at least 0.474 mol/m3. Furthermore, for a tube to form in the first place, the filaments need to overcome the initial resistive force, which is greater in magnitude than subsequent steady state load forces. Another salient requirement for successful protrusion formation is for the bundle to avoid buckling [68]. To elaborate on these ideas, we simulated an actin bundle polymerization against a sheet of planar membrane at a constant surface tension with fixed edge boundaries. The results are summarized in Fig. 2.5. Our simulations show that a bundle with 4 filaments cannot effectively protrude against the membrane because it is not strong enough to resist buckling (Fig. 2.5a/b). With more actin filaments comprising the bundle, the protrusion can form at sufficiently high G-Actin concentrations (Fig. 2.5c/d). When a finger-like protrusion is successfully formed, its rate of growth increases with both the number of filaments in the bundle, and the G-Actin concentration (Fig. 2.5e). We assumed in this work that the pointed ends of the filaments are fixed, so there is no overall retrograde flow [100] of actin filaments. We also assumed that the G-Actin concentration is uniform along a membrane tube. Under these assumptions, the dependencies of the growth rate on the number of filaments in the bundle and the G-Actin concentration are qualitatively in accord 33 Figure 2.5: Filament bundle deforming the membrane. Color codes for (a) and (c): Cyan: Mem- brane, Red: Actin filaments. (a) An actin bundle with 4 filaments cannot make a successful pro- trusion. (b) The maximum height of the membrane corresponding to (a) is shown. The growing filaments slightly deform the membrane in the initiation phase, subsequently becoming buckled and, hence, failing to form a tubular protrusion. (c) A bundle with 7 filaments form a finger-like membrane protrusion. (d) The protrusion height on the membrane corresponding to (c) is shown. Interestingly, bundles tend to buckle less when surrounded by the membrane, in agreement with a previous work [103]. (e) The growth rate for a nascent tubular protrusion increases almost lin- early as the G-Actin concentration increases, and also increases with the number of filaments in the bundle. with previous theoretical models [27, 68]. However, these assumptions allow the filopodia to grow at a constant speed indefinitely, which is unphysical in real filopodia. The existence of actin retrograde flow and the finite diffusion rate of G-Actin molecules lead to a steady state filopodial structure having a finite length and a dynamically generated G-Actin concentration gradient along the tube [69]. Furthermore, transportation of G-Actin by motor molecules, such as Myosin-X, may lead to a non-monotonic G-Actin concentration profile along the filopodium [101, 102]. These additional aspects of growth of mature filopodia are potentially amenable to simulations based on further elaboration of our new model. In real cells, the polymerization directions of filament bundles are not necessarily perpen- dicular to the membrane. At sufficiently high G-Actin concentrations, we anticipate that the 34 tubular protrusion is less likely to form when the filaments are inclined to the membrane surface, in particular, because those filaments would be more susceptible to buckling. To investigate how bundle’s orientation affects protrusion formation, we simulated a fila- ment bundle polymerizing towards a planar membrane sheet, where we varied the incident angle, α , which is the angle between the axis of the filament polymerization and the normal vector of the membrane. We also varied the load force at the tip of a stable protrusion, by modifying both the membrane tension and the bending rigidity simultaneously, while keeping their ratio con- stant, i.e. γ = χγ0, κ = χκ0 where γ0 and κ0 are the reference membrane tension and bending rigidity, and χ is a unitless parameter. In this way, the tubes formed would have the same radius R = √ κ0/2γ0 [68], but the load force at the tip for a stable protrusion changes as ‖ f‖ = χ2 f0. The results are demonstrated in Fig. 2.6. When the filament bundle is perpendicular to the mem- brane surface (α = 0), as the membrane rigidity increases (with increasing χ), the filaments first bend, and then eventually buckle, resulting in a protrusion failure. Furthermore, our simulations show that when filaments are not perpendicular to the membrane (α > 0), they easily buckle even under low membrane rigidities. The above-discussed results suggest that bundle’s orientation relative to the membrane is important in controlling the process of protrusion initiation. These insights help to explain the prior experimental observations showing that bundles growing within a vesicle can either be bent by the membrane and form ring-like structures, or protrude the membrane to form finger-like structures [104]. 35 Figure 2.6: Membrane protrusion behaviors are shown under different relative membrane rigidi- ties characterized by a unitless parameter χ (where a larger χ means larger membrane tension and bending modulus) and the incident angle α between the filament bundle and the membrane surface normal. All protrusions are made by a bundle comprised of 7 actin filaments. Color codes: Cyan: Membrane, Red: Actin filament, Green: α-actinin cross-linker. 2.4 Conclusions In this work we have addressed the problem of the interactions between the active matter and the plasma membrane, which are important for cell functions and have been a growing field of study. To study the rich behaviors arising from these mechanochemical interactions, we have developed a new force field for simulating the plasma membrane and its interaction with the ac- tive network. These algorithms were incorporated into MEYDAN, which is a broad simulation platform for active matter. The new membrane model enables simulations of rich cytoskeletal behaviors near the cell membrane, accounting for geometrical, mechanical and chemical details. On a single CPU, membrane-MEDYAN simulations can treat systems on the length scale up to several micrometers and time scales on the order of thousand seconds. Our software implementa- 36 tion allows to easily tune various model parameters, such as the elastic coefficients characterizing a membrane, and its initial shape. Our new model enables simulations of various cellular systems where membrane-cytoskeleton interactions play a salient role, such as lamellipodia, filopodia and dendritic spines. In the current work, we applied membrane-MEDYAN to simulate a membrane vesicle, as well as the membrane-cytoskeleton systems under a variety of conditions. Simulations of an empty membrane vesicle under hyperosmotic stress lead to membranes’ intricate crumpling. Simulations of vesicles containing actomyosin networks show that bundling of actin filaments is crucial for generating tubular membrane protrusions. We found that the initial growth rate of successful protrusions is determined by the thickness of the protruding actin bundle and the G-Actin concentrations at the growth tip. Furthermore, unsurprisingly, lowering membrane’s surface tension or bending modulus facilitates the tubulation process. Finally we showed that the inclination angle between the filaments and the membrane is also critically important for the formation of finger-like protrusions. However, softer membranes can significantly mitigate this constraint. One limitation of the current membrane model is the absence of membrane-associated proteins and membrane-cytoskeleton adhesions, which modulate both mechanical and chemical properties of the membrane-cytoskeleton system. Membrane based proteins, such as the BAR- domain proteins, can cause membrane remodeling by generating local membrane curvatures, being present in many cellular processes such as membrane trafficking and endocytosis [21, 22]. Membrane-cytoskeleton adhesions could increase the apparent rigidity of membrane attached to the actin cortex, which is important in cell blebbing [22, 58, 72]. They are also important in cre- ating membrane invaginations, which are important in initiating endocytosis [105]. Moreover, 37 chemical signaling via membrane receptors are known to regulate cytoskeleton behaviors [62]. Our current model can be naturally extended in future to enable studying of such intricate pro- cesses. 38 Chapter 3: Simulating membrane proteins and their dynamics 3.1 Introduction The cell has the membrane as its boundary, and uses the membrane to communicate with external environments. Cell membranes use the lipid bilayer as the “backbone”, while various types of proteins can attach to the lipid bilayer, shaping the mechanical and chemical proper- ties of the membranes. Membrane proteins are vital in cellular functions such as morphology control and mechanochemical signaling. For example, cadherin is a transmembrane protein that promotes and stabilizes cell-cell contact, while also regulating the contractility of cortical cy- toskeleton [106]. Curvature generating proteins, such as BIN-Amphiphysin-Rvs (BAR) domain proteins, can regulate membrane deformations, which are important, for example, in initiating endocytosis or dendritic spines [23,105,107,108]. The spatial distribution of membrane proteins can also regulate cell activities. As an example, B cell receptors (BCRs) form microclusters dur- ing B cell activation, mediating intracellular signaling [109,110]. Understanding how membrane protein dynamics affect the membrane-cytoskeleton dynamics helps unveiling their mechanisms in cellular sensing processes. The dynamics and spatial distribution of membrane proteins can be affected by protein in- termolecular interactions, protein-lipid interactions, as well as interactions with the cytoskeleton. Pairwise attractions among proteins could cluster the protein molecules, forming 2D liquid-liquid 39 phase separation (LLPS) on the membrane [16–19]. Protein-lipid interactions allow proteins such as BAR domains and Amphipathic lipid packing sensor (ALPS) motifs to exhibit higher affinity on specific local instantaneous curvature, by matching membrane geometry and sensing lipid packing defects, respectively, effectively acting as membrane curvature sensor [23, 111]. Computer simulations allow us to access the details in the complex interactions among the lipid bilayers, the membrane proteins and the cytoskeleton, thus making it possible to reveal the dy- namics of membrane proteins during important cellular processes. To achieve this, we base off this work of the cellular simulation platform, MEDYAN (Mechanochemical Dynamics of Active Networks), capable of simulating the cytoskeleton and its interactions the membrane (discussed in Chapter 2) [9, 112]. In this work, we first discuss the membrane protein dynamics with pairwise interactions and under external potential fields. Then we introduce our new method of simulating protein reaction- diffusion processes on the unstructured membrane surface mesh in the MEDYAN platform, while also discussing how the potential energy field can be accounted for by adjusting the reaction rates. The we discuss how this model can be applied to study the behavior of membrane curvature sensing and generating proteins. Simulations using the new models show that membrane protein diffusion can be hindered by energy barriers and confinement is stronger with higher energy barrier and slower diffusion coefficients. Apart from confinements, the phase separation can also occur in simulations if pairwise protein interactions are included. Furthermore, we will study how the curvature sensing proteins preferably dock on membrane pieces with various curvature values, and we show how curvature generation affect cytoskeleton’s ability to deform the membrane. 40 3.2 Methods 3.2.1 Thermodynamics of interacting proteins on the surface Consider N interacting particles living in Rd . Particle i lives is at position x(i). Particles have pairwise interactions that only depend on their distances: w(r) where r = |x(i)−x( j)|. Let X be a vector in RdN with X = (x1 (1),x 2 (1), ...,x d (1), ...,x d (N)). The overall interaction energy is W (X) = ∑ i< j w(|x(i)− x( j)|). (3.1) Let V (x) be a potential energy function that is the same for every particle. Then let U(X) = W (X)+∑iV (x(i)) to be the total potential energy of the system. Assume that the particles undergo overdamped dynamics, such that Ẋ i = 1 γ ∂iU +ξ (t), (3.2) where γ is the damping constant, ξ (t) is the thermal white noise satisfying Cξ (t) := 〈ξ (s)ξ (s+ t)〉 = 2Dδ (t), where D is the diffusion constant. The Einstein relation holds: βγD = 1. Let p(X , t) be the probability density of the system at state X given the time t, and satisfies ∫ V N p(X , t)dV N = 1. The Fokker-Planck equation is ∂t p(X , t) = ∑ i 1 γ ∂i((∂iU)p)+D∂i∂i p. (3.3) Eq. 3.3 is extremely hard to solve due to the high dimensionality of X . Therefore, we 41 assume that with a large enough number of particles, the particles form a density field n(x, t), such that we can separate one particle which interacts with this density field. Naively, the interaction energy as a result of adding this particle is V (x(i)) + ∫ V w(|x − x(i)|)n(x, t)dV . However, this energy may be inaccurate or even blow up if w(r) rapidly varies or becomes singular when r approaches 0, due to the inclusion of self-interaction. Instead, we can decompose w(r) into two parts: A short-range strong repulsive interaction and a long-range weaker interaction. For the strong repulsive interaction we can model it as an upper limit on local particle density n0. We will use µ(r) to represent the long-range weak interaction. At x= x(i), n(x, t) can be expanded to the second order as n(x, t)≈ n(x(i), t)+∑ j ∂ jn(x(i), t)(x j− x j (i))+ 1 2 ∑ jk ∂ j∂kn(x(i), t)(x j − x j (i))(x k − xk (i)). The function µ(|x− x(i)|) has spherical symmetry centered at x(i), so the particle energy from interaction can be written as R(x, t) := ∫ V µ(|x− x(i)|)n(x, t)dV ≈ n(x(i), t) ∫ V µ(|x − x(i)|)dV + 1 2 ∑ j ∂ j∂ jn(x(i), t) ∫ V µ(|x − x(i)|)(x j − x j (i)) 2dV = n(x(i), t) ∫ V µ(|x− x(i)|)dV + 1 2d ∆n(x(i), t) ∫ V µ(|x− x(i)|)|x− x(i)|2dV . Under this formulation, let p(x(i), t) be the probability density for particle to be at position x(i) at time t. The Fokker-Planck equation for this particle is ∂t p(x(i), t) = ∑ j 1 γ ∂ j [ ∂ j(R(x(i), t)+V (x(i)))p(x(i), t) ] +D∂ j∂ j p(x(i), t), (3.4) where R(x, t) = n(x, t)µ0 +∆n(x, t)µ2, µ0 = ∫ V µ(|x|)dV, µ2 = 1 2d ∫ V µ(|x|)|x|2dV, n(x, t)≤ n0,∀x, t, (3.5) 42 The formulation above can be extended to particle systems on a differential manifold as well. Instead of Rd , we have a d-dimensional manifold M . Locally one can use coordinates x ∈ Rd with diffeomorphism φ : Rd → M . Assume that M can be embedded in Rm where m ≥ d, with the inclusion map i : M → Rm. A Riemannian metric g can defined on M by pulling back the metric δi j in Rm. Now, with two points p(1), p(2) ∈ M , we can define their distance using either the geodesic distance or the Euclidean distance in the embedding space. We assume that the pairwise interaction function w(r) decays fast enough such that the difference between the two is negligible at relatively flat extrinsic curvature. Moving forward, we will use d(p(1), p(2)) to represent the geodesic distance between p(1) and p(2). With similar arguments, we may continue using v1 and v2 obtained in the flat space. With the volume form dV = √ det(g)dx1∧ ...∧dxd , we still require that ∫ M p(p(i), t)dV = 1. Eq. 3.4 now becomes ∂t p(p(i), t) = 1 γ div [ grad(R(p(i), t)+V (p(i)))p(p(i), t) ] +D∆p(p(i), t). (3.6) where ∆ = div◦grad is the Laplace operator. 3.2.2 Surface discretization and stochastic diffusion sampling Next, we will discretize the system for d = 2 and m = 3, that is, a 2D manifold embedded in R3. Let {q(i)} be the set of vertices of a triangulation M̂ of M , where the edges E are straight lines in R3 connecting neighboring vertices. Let y(i) be the coordinates of q(i) in a certain coordinate system. Let N1,v(q(i)) be the set of all 1-ring neighbor vertices of q(i) and N1,t(q(i)) be the set of triangles that has vertex q(i). Let t(q(i),q( j),q(k)) be the triangle formed by 3 unique 43 vertices. Now, we can discretize a function g : M → R supported by basis functions f j. f j : M̂ → R is defined on the triangulation, which evaluates to 1 at q( j) and 0 at all other vertices, and is linearly interpolated in each triangle in R3. Therefore, g(p) ≈ ∑ j g j f j(p) where p ∈ M . Let F be the vector space spanned by these basis functions. Eq. (3.6) can now be evaluated in the weak form using the inner product with the test function: 〈 f j,g〉 := ∫ M̂ f j(p)g(p)dV ≈∑k gk ∫ M̂ f j(p) fk(p)dV . Define M̃ jk := ∫ M̂ f j(p) fk(p)dV as the elements of the mass matrix, which is non zero only if j = k or { j,k} ∈ E. To simplify computation, mass lumping technique is applied to diagonalize the mass matrix [113]. After mass lumping, the new mass matrix can have non zero terms only on the diagonal: M j j = |C j|, where |C j| is the area around vertex q( j) and can be defined as a third of the 1-ring area ∑t∈N1,t(q( j)) A(t). Consequently, the weak form with mass lumping becomes 〈 f j,g〉= g j|C j|. With such finite element discretization, the first order derivatives (gradients) are well- defined on triangles, but the higher order derivatives are zero on the triangles and undefined on edges or vertices. To estimate higher order derivatives, we again use the weak formulation. Consider two functions g,h : M →R, and a second order derivative of form i := div(ggradh). If g is constant on M , then i = g∆h. Using the divergence theorem, we have 〈 f j, i〉= ∫ M̂ f j(p)div(ggradh)dV, = ∫ M̂ [ div( f jggradh)−g〈grad f j,gradh〉 ] dV = ∫ ∂M̂ f jg〈gradh, n̂〉dl − ∫ M̂ g〈grad f j,gradh〉dV, (3.7) where n̂ is the outward normal at the boundary of M̂ . With vanishing boundary conditions, the first term is zero. Therefore, 〈 f j, i〉=− ∫ M̂ g〈grad f j,gradh〉dV ≈−∑k,l gkhl ∫ M̂ fk〈grad f j,grad fl〉dV . 44 (a) q( j) q(n) αn βn (b) q( j) q(k) q(n1) q(n2) α γ β δ Figure 3.1: Geometric elements used in calculating a jk. (a) j = k. (b) j 6= k. Define a jk :=−∑l hl ∫ M̂ fk〈grad f j,grad fl〉dV , and then 〈 f j, i〉 ≈ −∑k a jkgk. When j = k, a j j = 1 6 ∑ q(n)∈N1,v(q( j)) (cotαn + cotβn)(h j −hn). (3.8) When j 6= k, a jk is non-zero only if {q( j),q(k)} ∈ E. In this case, a jk = 1 6 [ (h j −hk)(cotα + cotβ )+(h j −hn1)cotδ +(h j −hn2)cotγ ] . (3.9) The angles used in the equations above are illustrated in Fig. 3.1. With mass lumping, i j = −∑k a jkgk/|C j|. As a special case where g j = 1,∀ j, one has i = ∆h, and obtains the famous cotangent formula for the discrete Laplacian on surface mesh: (∆h) j = 1 2|C j| ∑q(n)∈N1,v(q( j)) (cotα+ cotβ )(h j −hn). Let S be a matrix such that ∑k S jkhk = |C j|(∆h) j,∀ j,{h j}, and S is also known as the “stiffness matrix” [113]. With the finite element discretization, Eq. 3.6 becomes |C j|∂t p j(t) = ∑ k 1 |Ck| ( − a jk γ +DS jk ) |Ck|pk, (3.10) 45 where a jk is defined with respect to functions g = p and h = R+V . For a small time period, the above equation can be applied to all particles in the system. From Eq. 3.10, the transition rate q jk for a particle to move from C j to Ck can be found as q jk = 1 |C j| ( − ak j γ +DSk j ) . (3.11) If the potential energy profile R+V is smooth, for a fine enough mesh, a jk scales with the mesh size, while S jk is invariant to the mesh size. Therefore, the physical transition rate q jk being positive puts a constraint on the maximum size of the mesh. In practice, we use a small enough mesh to ensure that the transition rate is positive most of the time. Considering rare cases where negative transition rate occur due to extraordinary local mesh shapes or a large energy gradient, we manually set the transition rate to zero. On the other hand, the transition rate is also set to zero if the transition causes the target cell to violate the nk ≤ n0 constraint. 3.2.3 Membrane-related reactions Apart from membrane species diffusions, we also incorporate in our model the membrane- related reactions (MRRs), which are the reactions happening at the membrane surface, but in- cluding at least one 3D diffusing species. Specifically, we want to model the reactions of form Mr1 +Mr2 + ...+Cr1 +Cr2 + ... M−→ Mp1 +Mp2 + ...+Cp1 +Cp2 + ... (3.12) where Mri, Mpi are 2D diffusing membrane species as reactants and products, respectively, and Cri, Cpi are 3D diffusing species as reactants and products, respectively. Assuming that the law 46 of mass action applies to the membrane domain, and given a small patch of membrane with area A, the membrane species and nearby 3D diffusing species are uniformly distributed, then the propensity of such a reaction is related to the rate constant r as p = rA∏ i [Mri]∏ i [Cri], (3.13) where the operator [·] returns the area density for membrane species and the volume density for 3D diffusing species. Some common cellular events can be described using MRRs: Adsorption / membrane binding C M−→ M Desorption / membrane unbinding M M−→C Membrane species activator M+C1 M−→ M+C2 Membrane activator (implicit) C1 M−→C2 Membrane species dimerization M1 +M1 M−→ M2 Similar to membrane diffusion processes, the potential energy field on the membrane also affects reaction rates. For example, let the reaction rates of a reversible adsorption/desorption reaction be rad and rde, respectively, and their ratio encodes the free energy change between the bound and the unbound states. If the protein is subject to extra potential fields on the membrane, we will adjust the desorption rate to make up for that: rde,eff = rdeeβV . (3.14) 47 3.2.4 Free energy of the particle system Following the derivation in Ref. 114, the total free energy can be found by E = ∫ M ( − 1 β φ(n(p))− 1 4 ∫ M µ(d(p, p′)) [ n(p)−n(p′) ]2 dS′+V (p)n ) dS, (3.15) where φ(n) = n ln [n0−n n ] + n− 1 2β µ0n2. Assume that in the neighborhood of p, one can use a geodesic frame x̃ to represent surface positions, such that it is isometric to δi j in R3 at p, and close to being isometric near p. Then one can expand the second term above, such that E ≈ ∫ M − 1 β φ(n)− 1 4 ∫ M µ(|x̃− x̃′|) [ ∑ i ∂x̃in(x̃)(x̃− x̃′)i ]2 dS′+V n dS = ∫ M ( − 1 β φ(n)− 1 4 〈gradn,gradn〉 ∫ M µ(|x̃− x̃′|) [ (x̃− x̃′)‖ ]2 dS′+V n ) dS = ∫ M ( − 1 β φ(n)− µ2 2 〈gradn,gradn〉+V n ) dS, (3.16) where (x̃− x̃′)‖ is the unit vector along the gradn direction. In MEDYAN, we select the enthalpic part of the energy above, which enters energy mini- mization process. The entropic energy minimization is implicitly done via the stochastic diffusion- reaction sampling mentioned above, so that the dynamics of protein distribution equilibration is retained. 48 3.2.5 Membrane curvature sensing and generation model A protein binding to the membrane locally with curvature H results in free energy change ∆E = Eb +VH(n,H), (3.17) where Eb < 0 is the fixed portion of binding energy, and VH(n,H) is the energy dependent on local protein concentrations and the curvature. Various computational models were developed to provide an estimate to VH(n,H) [23]. Two simplest examples are the spontaneous curvature model and the curvature mismatch model [115, 116]. In the spontaneous curvature model, the bending energy density of the protein-bound membrane surface is still in the Helfrich form: ebend = 2κ(H −H0) 2, where the bending rigidity κ is a constant, but the spontaneous curvature H0 depends on the membrane protein’s mechanical characteristics and its local concentration. On the other hand, the curvature mismatch model treats the lipid bilayer and the proteins separately, with ebend = 2κl(H −H0) 2 + 2φpκp(H −Hp) 2, where κl and κp are the bending rigidities of the lipid bilayer and the protein, respectively, Hp is the spontaneous curvature of the membrane bound protein, and φp is the fraction of area occupied by the protein. If the area of individual membrane protein is Ap, then φp = nAp. Hybrid models where both the bending rigidity and the spontaneous curvature depend on protein concentration are used as well [29]. In this work, we adopt the following bending energy density expression [117] ebend = 2κ(H −H0φp) 2, (3.18) 49 where κ = ( 1−φp κl + φp κl +κp )−1 H0 = Hp κp κl +κp (3.19) When φp = 0, Eq. 3.18 reduces to the simple Helfrich form for lipid bilayer with zero equilibrium curvature. For this reason, the constant used in Eq. 3.18 is scaled from 1/2 to 2 to match the Helfrich form conventions in this work. The appendix provides more discussions on this energy formula. In the discretized mesh model, VH(n,H) at each vertex can be obtained from the difference in bending energy before and after adding an additional molecule to the cell around this vertex, ie VH(n,H) = |C |[ebend(n+ |C |−1)− ebend(n)]. If a cell already contains many such molecules (where n|C | � 1), one can approximate VH(n,H) using VH(n,H)≈ ∂nebend(n). 3.3 Results 3.3.1 Energy barriers acting as protein confinements decrease diffusivity With our new membrane-protein model, we simulate membrane protein diffusions with diffusion barriers, where we also benchmark the appropriate mesh size under specific potential energy gradients. In reality, the surface diffusion barriers can arise due to dense clustered mem- brane proteins [67] or nearby actomyosin cortical networks [57]. We initialized a piece of membrane of 1500nm× 1500nm parallel to the xy plane. The membrane species M can diffuse on the membrane under a potential energy field. The potential energy field is a constant along the y direction and is a Gaussian function in the x direction with center at x = 750nm and width 400 nm, which represents a linear diffusion barrier parallel to the 50 y axis. The maximum value of the potential energy is varied from 0 to 4 kBT (Fig. 3.2ab). The membrane domain is triangulated, and the characteristic mesh size is controlled during both mesh generation and adaptation processes, with a value in 20 – 300 nm. We added 22500 copies of M uniformly at region x < 500nm at initialization (t = 0). Then, we start the simulation and let the species diffuse in the surface. We track the number of molecules on the right half of the plane (x > 750nm) at every frame, which starts at 0 and goes toward the equilibrium value over time. At equilibrium, due to symmetry, the average molecule count on the right should be a half of the initialized copy number. Unless otherwise specified, all parameters used are listed in Table. 3.1. Fig. 3.2c show trajectories of molecule counts on the right under various barrier heights. Notice that when the height is 0, the potential energy is zero everywhere on the surface, meaning that the proteins undergo free diffusion. The equilibration rate is slower with higher barriers, resembling a slower state transition rate under higher activation energy. With a constant barrier height, the equilibration rate increases with increasing diffusing coefficient of M (Fig. 3.2d). In Fig. 3.2e, we also show that the equilibrium can be reached much faster if the protein can unbind from the membrane and diffuse in 3D in addition to in-membrane diffusion, circum- venting the energy barrier in this pathway. In a triangulated meshwork, the transition rate in Eq. 3.11 should be non-negative, which sets an upper limit to the mesh size given potential energy gradients. As discussed in Appendix B.1 and Fig. B.2, the upper limit of permissible mesh sizes decreases with the height of the energy function. 51 Table 3.1: Membrane diffusion barrier simulation parameters. Description Value Default 2D diffusing coefficient (D2D) 5×106 nm2/s a Default 3D diffusing coefficient (D3D) 1×108 nm2/s b Protein binding rate (rad) 6000 nm/s c Protein unbinding rate (rde) 10 s−1 d aFrom Ref. 118, typical diffusion rates for membrane proteins with radius under 5 nm are between 3×106 nm2/s and 6×106 nm2/s. bEstimated using Stokes-Einstein equation D3D = kBT/6πηr, where η is the dynamic viscosity of water (∼ 1×10−6 pNs/nm2) and r is the protein radius (∼ 2.5nm) [119]. cProtein binding and unbinding rates are chosen such that in equilibrium, ∼ 22500 molecules are bound on the membrane with ∼ 40000 molecules in total, and that this equilibrium can be reached before 0.4 s. dSee footnote c. Figure 3.2: Membrane species diffusion with a barrier. (a) On a piece of square membrane, the energy barrier is initialized as a Gaussian function of various heights with 400 nm width along the x direction and is uniform along the y direction. The triangular mesh discretization is shown in gray. (b) The energy function described in (a) in the x direction. The proteins initialized on the left must overcome the energy barrier to reach the right. (c) Fraction of equilibrium protein count on the right as a function of time, varying barrier heights. Equilibration is slower with higher barrier energy. Diffusion coefficient: 5×106 nm2/s. (d) Fraction of equilibrium protein count on the right as a function of time, varying diffusion coefficients. Equilibration is faster with higher diffusion coefficients. Barrier height: kBT . (e) If the protein can unbind from the membrane and has a higher 3D diffusion rate, it can overcome the barrier easily, and greatly accelerates the equilibration process. 52 3.3.2 Pairwise protein interactions promote LLPS To incorporate protein pairwise interactions in our simulations, we need a pairwise at- tractive potential µ(r) that µ0 and µ2 defined in Eq. 3.5 are finite. If one normalizes the pairwise interaction function resulting in a 2D distribution f (x) := µ(|x|)/µ0, its standard de- viation σ characterizes the range of the interaction. σ can be represented by µ0 and µ2 as σ2 = ∫ V f (x)|x|2dV = (1/µ0) ∫ V µ(|x|)|x|2dV = 4µ2/µ0. For simplicity, we model µ(r) as a Gaussian function with µ(r) = −hexp(−r2/σ2), where h ≥ 0 is the height of this energy func- tion. In this case, µ0 =−πhσ2 and µ2 =−πhσ4/4. We note that if the attractive potential takes the form of a power function µ(r) =−hrp where p <−1, µ0 necessarily diverges when h > 0. We initialized a square membrane of size 400nm×400nm, and uniformly initialized 1500 molecules bound to the membrane. Then, we ran chemistry simulations, allowing those proteins to diffuse. The simulation parameters are displayed in Table 3.2. The simulation lasts 0.5 s and snapshots are captured every 0.01 s starting at 0.11 s. When there is no protein attractive interaction, the protein shows gaseous behavior. As the protein interactions become stronger, protein condensates that span multiple membrane cell volumes are formed, showing a liquid phase separation behavior. Fig. 3.3b and c shows two snapshots when the attractive interaction is weak (h = 0.01kBT ) and strong (h = 0.21kBT ), respectively. The equilibrium area fraction distribution is collected at a variety of energy heights and is displayed in Fig. 3.3a. The phase separation starts at h ∼ 0.09kBT , where the bifurcation showing phase separation starts to appear beyond that. In Appendix B.2, we show that the initial protein distribution has no observable effect on the final phase separation results. In LLPS-manifesting simulations, the range of mesh sizes faces stronger restrictions. It 53 Table 3.2: Protein LLPS simulation parameters. Description Value Diffusing coefficient 106 nm2/s a Protein area (Ap) 19.6 nm2 b Interaction range (σ ) 18 nm c Target mesh size 12 nm aThe same order of magnitude as described in footnote a in Table 3.1. bEstimated using protein with a circular membrane projection of radius 2.5 nm. cWe choose pairwise interaction ranges with the order of magnitude ∼ 10nm. This is backed by the experimental results in Ref. 120, where the minimum pairwise interaction energy between ATP synthase c-rings happens at a distance of 10.3 nm and drops below 1kBT for distances larger than 12.5 nm. should not be greater than the protein interaction range σ , otherwise high potential energy gra- dient (mentioned in the previous section) will emerge on the cluster boundaries, leading to er- roneous clustering behaviors. An example is presented in Appendix B.1 and Fig. B.3. On the other hand, the mesh size should still be greater than the protein size. Consider a hypothetical membrane mesh where every cell has size 1.5Ap. Since each cell contains an integral number of proteins, a cell can never contain 2 such proteins simultaneously and the remaining 0.5Ap can never be utilized by the proteins, which effectively increases the protein area by 50%. When the size is larger, such discrepancy (1−b|C |/ApcAp) can be reduced. Due to the range restrictions of mesh sizes, this model is more useful for small proteins with long-range attractive pairwise interactions. 3.3.3 Concentration sorting by local curvature To simulate protein curvature sensing behavior, we use Eq. 3.18 to determine the surface potential energy. We initialized membrane vesicles with various radii while keeping their shapes fixed after initialization. Proteins were added to the system as 3D diffusing species, that can bind to and unbind from the membrane surface (Fig. 3.4a). The effective unbinding rate depends on 54 Figure 3.3: Membrane protein phase separations. (a) A snapshot for the protein distributing on the membrane in a gas phase. h = 0.01kBT . (b) A snapshot of the protein with liquid conden- sations on the membrane. h = 0.21kBT . (c) The probability distribution of area fraction (φp) for different interaction strengths (h). When h = 0.21kBT , the distribution becomes bimodal, dis- playing phase separation behavior. (d) The area concentration profile among membrane cells, at various protein attractive interaction strengths quantified by its height. The brighter color indi- cates a higher probability in the distribution. When the pairwise interaction is weak, the diffusion takes dominance and the system is in a gas phase. When the height of the energy function goes above ∼ 0.09kBT in