<<

Revised manuscript Directed Growth of Biomimetic 1 Microcompartments 2 3 4 5 Directed Growth of Biomimetic Microcompartments 6 7 8 9 1 Ivan Ivanov1*, Rafael B. Lira2, T-Y. Dora Tang3, Titus Franzmann3, Adam Klosin3, Lucas 10 11 4 3 4 2 12 2 Caire da Silva , Anthony Hyman , Katharina Landfester , Reinhard Lipowsky , Kai 13 14 3 Sundmacher1, Rumiana Dimova2* 15 16 17 4 1Max Planck Institute for Dynamics of Complex Technical Systems, Process Systems Engineering, 18 19 20 5 Sandtorstrasse 1, 39106 Magdeburg, Germany 21 22 2 23 6 Max Planck Institute of and Interfaces, Theory and Bio-Systems, Science Park Golm, 24 25 7 14424 Potsdam, Germany 26 27 28 8 3Max Planck Institute of Molecular Biology and Genetics, Organization of Cytoplasm & 29 30 31 9 Dynamic Protocellular Systems, Pfotenhauerstrasse 108, 01307 Dresden, Germany 32 33 4 34 10 Max Planck Institute for Research, Physical Chemistry of , Ackermannweg 10, 35 36 11 55128 Mainz, Germany 37 38 39 40 12 * Correspondence: 41 42 13 Corresponding Authors 43 44 14 [email protected], [email protected] 45 46 47 48 15 Keywords: growth, vesicles, -rich droplets, , polymersomes, fusion, fatty 49 50 16 acids 51 52 17 53 54 18 Abstract 55 56 57 19 Contemporary biological cells are sophisticated and highly compartmentalized. 58 59 20 Compartmentalization is an essential principle of prebiotic as well and a key feature in bottom-up 60 21 synthetic biology research. In this review, we discuss the dynamic growth of compartments as an 61 62 63 64 1

65 Directed Growth of Biomimetic 1 Microcompartments 2 3 22 essential prerequisite for enabling self-reproduction as a fundamental life process. We focus on 4 5 23 micron-sized compartments due to their cellular dimensions. Two types of compartments are 6 7 24 considered, membraneless droplets and membrane-bound microcompartments. We review growth 8 9 25 mechanisms of aqueous droplets such as protein (condensates) or -rich droplets 10 11 26 (aqueous two phase systems) and coacervates, for which growth occurs via Ostwald ripening or 12 27 coalescence. For membrane-bound compartments, we consider vesicles, composed of fatty acids, 13 14 28 lipids or polymers, where directed growth can occur via fusion or uptake of material from the 15 16 29 surrounding. The development of novel approaches for growth of biomimetic microcompartments 17 18 30 can eventually be utilized to construct new synthetic cells. 19 20 31 21 22 23 24 32 1 Introduction 25 26 27 33 Despite the long and ongoing debate about the essence of life, there is a widespread consensus about 28 29 30 34 some common features of all living systems. While some phenomena, such as metabolism, remain 31 32 35 arguable benchmarks (and place in the twilight zone), it is widely recognized that the 33 34 36 formation of micro- and nanocompartments is an essential ingredient of all forms of life. This spatial 35 36 37 37 constraint serves numerous purposes, including the segregation and protection from the environment 38 39 38 (to allow for individuality and maintenance), the establishment of gradients (to enable and make use 40 41 42 39 of out-of-equilibrium conditions), and the role of two- and three-dimensional confinement for self- 43 44 40 organization phenomena, all of which serve to overcome the overall “dilution problem”. In order to 45 46 47 41 fulfil another cornerstone of life – proliferation – compartments also need to change with time by 48 49 42 fusion, growth, and division. In particular, the dynamic growth of compartments is an essential 50 51 52 43 prerequisite for enabling self-reproduction as a fundamental life process, both in simplistic systems 53 54 44 such as droplets or fatty-acid based vesicles, as well as for lipid vesicle compartments with 55 56 45 membranes that resemble the biomembranes of today’s cells. In this context, we argue that growth 57 58 59 46 deserves more attention, not only because growth precedes division but also because of the difficulty 60 61 47 to realize growth compared to division, especially in the case of lipid vesicles, where budding and 62 63 64 2

65 Directed Growth of Biomimetic 1 Microcompartments 2 3 48 division has been observed in response to various factors This growth aspect has also been 4 5 6 49 recognized in basic theoretical models of living systems such as Ganti’s chemoton [1, 2], for which 7 8 50 the increase of membrane area (referred to as membrane formation) was postulated to be one of three 9 10 11 51 subsystems, characterizing living entities from a chemical viewpoint. The autopoietic theory was also 12 13 52 centered on the membrane but focused on self-maintenance rather than on (self-)reproduction [3, 4]. 14 15 53 However, such a self-maintaining system could also enter a self-reproduction mode, manifested by 16 17 18 54 growth, if homeostatic misbalance leads to excess membrane formation as shown in a thought 19 20 55 experiment by Luisi [3]. 21 22 23 24 56 In this paper, we review the existing approaches for growth of compartments in the context of 25 26 57 bottom-up synthetic biology and protobiology. We consider mainly micron-sized compartments due 27 28 29 58 to their characteristic cellular dimensions; this feature also ensures that the area and curvature of the 30 31 59 interface or the bounding membrane has less influence on the enclosed solution. Microcompartments 32 33 34 60 of various origins and chemistries have been used as models, and many studies have 35 36 61 addressed growth and division simultaneously in an ambitious effort to mimic self-reproduction. 37 38 62 Here, we focus on growth and divide the literature examples into two categories: membraneless and 39 40 41 63 membrane-bound microcompartments, whereby we review aqueous droplets (such as protein- or 42 43 64 macromolecule-rich droplets and coacervates) on the one side, and vesicles, composed of fatty acids, 44 45 46 65 lipids or polymers, on the other. This is by no means a universal and comprehensive classification of 47 48 66 compartment types in nature1, but rather an attempt to classify reported examples. 49 50 51 52 53 54 55 56 57 58 59 60 1 61 We conceptually include examples of fatty acid vesicles as presumable predecessors, as well as vesicles made of 62 synthetic chemicals as prospective successors of today’s cells. 63 64 3

65 Directed Growth of Biomimetic 1 Microcompartments 2 3 67 2 Coacervates and protein-rich droplets – volume and area grow simultaneously 4 5 6 7 68 In the 1920’s Oparin hypothesized that membrane-free liquid-like droplets formed via complex 8 9 69 coacervation could have been precursors to modern cells [5, 6]. Extensive research has demonstrated 10 11 12 70 that small molecules and oppositely charged polymers can phase separate to form structures with 13 14 71 emerging liquid-like properties. More recently, it was discovered that certain types of 15 16 72 undergo intracellular phase separation to form membrane-free/membraneless liquid-like protein-rich 17 18 19 73 droplets. Both of these membraneless microcompartments represent interesting systems to 20 21 74 circumvent the coupling between volume and surface area, and their austerity in comparison to 22 23 24 75 membrane-bound microcompartments has qualified them as useful models to demonstrate 25 26 76 compartment formation and growth. The relevance of membraneless compartments to bottom-up 27 28 29 77 synthetic biology, regardless of their chemical composition, remains undisputed, since they 30 31 78 successfully represent or mimic certain biological aspects. Thus, phase-separated droplets represent 32 33 34 79 useful models of self-organized entities, independent of the natural or man-made origin of the 35 36 80 . Form a protobiology perspective, coacervates made of synthetic polymers are 37 38 81 logically excluded as potential ancestors of today’s cells, but such coacervates have played an 39 40 41 82 important role in the history of , starting with the first experiments on gum arabic and 42 43 83 gelatin. 44 45 46 47 84 2.1 Protein-rich droplets 48 49 50 85 Compartmentalization is a hallmark of the eukaryotic cell. It enables cells to spatially separate their 51 52 53 86 complex biochemistry into microreactors. Well-known examples of compartments are the 54 55 87 and the nucleus, both of which are surrounded by lipid bilayers. However, in recent 56 57 58 88 years it has become clear that the eukaryotic cytoplasm is further organized by compartments that 59 60 89 lack membranes. These compartments have been termed biomolecular condensates, also known as 61 62 63 64 4

65 Directed Growth of Biomimetic 1 Microcompartments 2 3 90 membraneless organelles [7]. Intracellular condensates form by the process of liquid-liquid phase 4 5 6 91 separation [8-13]. Phase separation is a highly cooperative and concentration-dependent process: 7 8 92 above a certain saturation concentration, a solution becomes unstable and demixes to form two or 9 10 11 93 more distinct co-existing aqueous phases, a protein-dense and a protein-poor phase that continuously 12 13 94 exchange molecules and maintain a steep concentration gradient across their interfaces. Recent work 14 15 95 suggests, that the formation of biomolecular condensates is primarily driven by RNA and proteins 16 17 18 96 that exhibit a large degree of intrinsic disorder and/or multivalency. Among the growing list of 19 20 97 proteins that undergo liquid-like phase separation are the well-studied P- proteins LAF1 and 21 22 23 98 PGL-3 [14, 15], the proteins FUS and hnRNPA1 [16-18], the centrosome protein SPD- 24 25 99 5 [19], as well as the nucleolar proteins Fib1 and Npm1 that form immiscible phases to organize the 26 27 28100 [20]. When expressed in a test tube, these proteins readily phase-separate to form two- 29 30101 phase systems at physiological concentrations. These protein-rich droplets grow by taking up 31 32102 material from the environment (Ostwald ripening) or via coalescence of two smaller droplets to form 33 34 35103 a larger one; see Figure 1. The condensates exhibit liquid-like characteristics, such that they can 36 37104 coalesce, wet surfaces and deform under shear stress. The dynamics of these processes is governed 38 39 40105 by internal viscosity and surface tension [8, 14, 21]. Importantly, condensates display selectivity, and 41 42106 client molecules such as interacting proteins or RNA will partition into the condensates. 43 44 45 46 47 48 49 50 51 52 53 54 55 56107 57 58 59108 Figure 1. (A) Protein-rich droplets can grow by the uptake of material from the surrounding and at 60109 the expense of smaller droplets (Ostwald ripening). (B) Still fluorescence images from a time-lapse 61 62 63 64 5

65 Directed Growth of Biomimetic 1 Microcompartments 2 3110 movie of a Hela cell expressing the YFP-labelled disordered region of the RNA helicase DDX4 4111 (unpublished data). The disordered region forms an intracellular liquid-like droplet. With time, the 5 6112 bigger droplet grows, whereas the smaller droplets shrink and eventually disappear. (C) 7113 Quantification of the bigger droplet area shown in (B) as a function of time (unpublished data). (D) 8114 Growth of liquid-like protein droplets can also occur via coalescence upon encounter of two or more 9115 droplets that merge to form a bigger one. (E) Still bright field images from a time-lapse movie of 10 11116 protein droplets formed from the protein Sup35 [22]. The protein droplet was held with an 12117 optical tweezers and brought in the vicinity of other droplets, which coalesce leading to growth with 13118 time. (F) Quantification of the droplet area shown in (E) as a function of time. The step-wise increase 14119 in area is indicative of coalescence. 15 16120 Recent work has identified that the saturation concentration of the FET family of proteins (FUS- 17 18 19121 EWSR1-TAF15 protein family) is quantitatively determined by the associative cation-pi interactions 20 21122 between segregated tyrosine and arginine sticker residues [23]. Other types of interactions, such as 22 23 24123 pi-pi and electrostatics between charged residues, play important roles as well [24]. For instance, the 25 26124 disordered region of the condensate protein helicase DDX4 displays a great degree of segregated 27 28 29125 charges [25]. Importantly, weak multivalent interactions underlie the nucleation process [11]. 30 31 32126 In the context of the cell, these systems are driven away from equilibrium, e.g. by chemical reactions. 33 34 35127 However, in the test tube, droplet growth continues until the thermodynamic equilibrium has been 36 37128 reached. Accordingly, mechanisms must exist that regulate the nucleation as well as the size and the 38 39129 mechanical properties. Liquid-like condensates can also undergo liquid-to-solid transitions [18, 22, 40 41 42130 26, 27]. For FET family proteins, the composition of interspacing residues (so-called spacers) that 43 44131 segregates the stickers plays an important role in modulating the interaction strength and thus the 45 46 47132 liquid-to-solid transition [23]. The liquid-to-solid transition of the centrosome forming protein SPD-5 48 49133 may be mediated by coiled-coil domains to establish a force-resistant meshwork for microtubule- 50 51 52134 mediated segregation [19]. Coupling phase separation to gelation may provide an ideal 53 54135 size-determining mechanism. Moreover, coupling the phase behavior to biochemical reactions that 55 56136 drive the system away from equilibrium establishes control over nucleation, droplet growth and 57 58 59137 under certain circumstances may even provide a mechanism for droplet division [28, 29]. Phase 60 61138 separation is exquisitely sensitive to environmental changes, including temperature, ionic strength or 62 63 64 6

65 Directed Growth of Biomimetic 1 Microcompartments 2 3139 pH, parameters that also fluctuate within cells [30, 31]. Cells have harnessed the phenomenon of 4 5 6140 phase separation to orchestrate complex processes and evolved precise mechanisms to control 7 8141 nucleation of biomolecular condensates. Taken together, it is now possible to predict the saturation 9 10 11142 concentration, as well as the material constants for minimal proteinaceous systems. Linking synthetic 12 13143 phase separation units with enzyme functions provides an opportunity to organize and polarize 14 15144 biochemical reactions within a synthetic cell. 16 17 18 19145 2.2. Polymer-rich droplets 20 21 22146 microdroplets are generated from associative liquid-liquid phase separation 23 24 25147 (coacervation) processes between oppositely charged polymers or small highly charged molecules. 26 27148 They form from synthetic polymers (PDDA, polyethylemine, etc.) [32, 33], as well as from 28 29 30149 biologically relevant molecules including proteins (BSA) [34], polypeptides (polylysine, 31 32150 polyarginine) [35, 36], nucleic acids (DNA, RNA) [33, 37, 38], and co-factors (ATP, FAD, NAD) 33 34 35151 [39]. In comparison, dissociative liquid-liquid phase separation into two aqueous phases can be 36 37152 triggered in solutions of neutral macromolecules such as polyethylene glycol (PEG) and dextran, and 38 39153 more recently, this process was also established in the closed compartment of lipid vesicles [40, 41]. 40 41 42154 It is commonly considered that the general mechanism of phase separation occurs via two steps: an 43 44155 enthalpic contribution to the free energy from the electrostatic interaction between the molecules (in 45 46 47156 the case of charged polymers), which draws the molecules towards each other, and an entropic 48 49157 driving force from the rearrangement of ions and water leading to a lowering of the Gibbs free energy 50 51 52158 and the formation of membraneless, chemically enriched microdroplets. It has been proposed that this 53 54159 mechanism is analogous to liquid-liquid phase separation in biology [42, 43]. The surface tension of 55 56160 coacervates is low, between 1 µN/m – 1 mN/m [44], and even lower for aqueous two-phase systems 57 58 59161 of neutral polymers [45]. The droplets will grow in size over time via coalescence events [46]. The 60 61 62 63 64 7

65 Directed Growth of Biomimetic 1 Microcompartments 2 3162 rate of coalescence and therefore growth can be tuned by the overall charge ratios of the coacervate 4 5 6163 components. 7 8 9164 Studies have shown that polymer-rich microdroplets will selectively partition a range of different 10 11 12165 molecules leading to molecular localization and further chemical enrichment [39, 47-49]. Moreover, 13 14166 enzymes, which partition into the microdroplets maintain activity within the highly charged and 15 16167 crowded interior [50, 51]. Enzymes capable of phosphorylation and dephosphorylation have led to 17 18 19168 cycles of growth and degradation of the coacervate droplets [37, 52]. In addition, growth and decay 20 21169 of coacervate droplets can be instigated by switching the pH of the solution by bubbling carbon 22 23 24170 dioxide and ammonia through the dispersion of coacervate droplets [39]. These results indicate that 25 26171 the physical parameters such as molar ratio of polymers or pH and temperature of coacervate droplets 27 28 29172 can be used to tune and drive droplet growth and disassembly. Additionally, the ability of 30 31173 membraneless coacervate droplets to partition and support enzyme reactions could suggest that these 32 33 34174 systems could be plausible models for predicted growth and division cycles driven by flux of 35 36175 molecules across the interface [29]. 37 38 39176 3 Fatty acid vesicles as primitive protocells – membrane formation due to fast equilibrium 40 41 42 43177 The vesicular (membrane) systems represent another compartment type with the same structural 44 45178 features as living cells. Worth noting with respect to the field of minimal cells, and growth in 46 47 48179 particular, are the pioneering works of Luisi and Szostak [53, 54], which mainly relied on fatty acid 49 50180 micelles and vesicles. The latter have been investigated extensively as protocell models because fatty 51 52 53181 acids have been proposed to be prebiotically relevant due to their structural simplicity compared to 54 55182 phospholipids [55]. Therefore, the spontaneous uptake of fatty acids into preformed vesicles has 56 57 58183 often been considered as a primitive growth mechanism. As far as structural simplicity is concerned, 59 60184 the emphasis on fatty acids is analogous to the RNA hypothesis (single-/double-stranded vs. one or 61 62 63 64 8

65 Directed Growth of Biomimetic 1 Microcompartments 2 3185 two hydrophobic tails) but the quest for the origin of life is intrinsically speculative and we will 4 5 6186 refrain from statements about the evolutionary relevance of these compartments. Regardless of this 7 8187 open question, fatty acid vesicles and fatty acids have served as rewarding models for mimicking life 9 10 11188 processes, including growth. 12 13 14189 We also note that the majority of studies with fatty acids employed small (~30-50 nm) and large 15 16190 (~100-200 nm) unilamellar vesicles – below the limit of optical resolution – which poses the question 17 18 19191 whether those are relevant protocell models from another perspective, namely their small size. A 20 21192 workshop dedicated to defining the minimal compartment size of life, related to possibilities for 22 23 24193 extraterrestrial life and nanobacteria, dealt with this question 20 years ago, see proceedings [56]. The 25 26194 biologists view regarding the threshold size of primitive compartments was set to diameters of 27 28 29195 250±50 nm determined by a requirement for the compartment to accommodate minimum amount of 30 31196 necessary proteins, and ribosomes. Others views were more speculative suggesting ⁓50 nm as 32 33 34197 the minimal size required to sustain potential forms of primordial life. The latter notion was 35 36198 supported by reports on enhanced protein expression in 100 nm liposomes [57]. Considered from a 37 38199 simplistic chemoton perspective, smaller size implies higher surface-to-volume ratio, which is 39 40 41200 associated with enhanced inward flux (proportional to the surface) of membrane precursors and other 42 43201 metabolites and thus could be considered as beneficial. However, other factors such as packing, 44 45 46202 curvature, etc., could result in micrometer size (as in modern cells) optimal for a specific functions. 47 48203 In this review, we focus on biomimetic microcompartments, while we do not aim to set the size 49 50 51204 limits of protocells. However, we include reports on nanocompartments as well, due to their 52 53205 historical and mechanistic significance. Furthermore, microcompartments such as giant unilamellar 54 55206 vesicles (GUVs) [58-61] have proven to be a practical system, whose properties, growth and 56 57 58207 response to external factors can be monitored and manipulated directly under a microscope. 59 60 61 62 63 64 9

65 Directed Growth of Biomimetic 1 Microcompartments 2 3208 Compared to phospholipids, fatty acids exhibit a very quick exchange between the membrane leaflets 4 5 6209 as well as faster exchange kinetics between the vesicle and the solution. In addition, stabilizing head 7 8210 group interactions are more important, which endows distinct properties, such as growth and self- 9 10 11211 reproduction of vesicles [62-64]. Due to their single-chain structure, the concentration of monomers 12 13212 in equilibrium with vesicles is significantly higher than for phospholipids, which enables fast flip- 14 15213 flop and exchange of molecules; Figure 2A. These beneficial properties have qualified fatty acids as 16 17 18214 a prominent system for studying protocells, including the demonstration of membrane growth [62]. A 19 20215 notable phenomenon, observed during the growth of oleate vesicles, was the so-called matrix effect, 21 22 23216 which exemplified itself when a seed of preformed vesicles was added to the micellar oleate solution, 24 25217 resulting in a narrow size distribution, corresponding to the size of the seed [65]. The growth of 26 27 28218 oleate vesicles was further investigated in detail, which lead to kinetic [66] and molecular dynamics 29 30219 [67] models. The growth process was also demonstrated under flow [68], whereby filamentous 31 32220 microcompartments were formed, and also employed as a mechanism for activation [69] to 33 34 35221 manifest a form of homeostasis. In addition, the filamentous growth of multilamellar oleate vesicles 36 37222 (⁓4 µm) was used to mimic the full proliferation cycle of growth and division [70] and the chemistry 38 39 40223 was extended to other surface-active molecules beyond oleate [71]. 41 42 43 44224 While fatty acids and surfactants are efficiently incorporated into existing membranes due to their 45 46225 optimal physicochemical properties, phospholipids – the membrane constituents of modern living 47 48 49226 cells – form structures, which are stable over a wider range of concentrations (phospholipids exhibit 50 51227 lower critical micelle concentration, CMC, compared to fatty acids). With respect to this distinction, 52 53228 the addition of oleate to preformed phospholipid vesicles was used in other attempts to mimic self- 54 55 56229 reproduction as an intermediate approach. Growth and division has been shown and investigated in 57 58230 detail in several examples of fatty acid incorporation into smaller liposomes [72-74] and the method 59 60 61231 has been later on scaled up to GUVs [75, 76]. In a similar fashion, the addition of detergents and 62 63 64 10

65 Directed Growth of Biomimetic 1 Microcompartments 2 3232 other water-soluble membrane active molecules at low concentration leads to their intercalation into 4 5 6233 the bilayer, which in turn could lead to an increase in the membrane area. In GUVs this increase 7 8234 depends on whether the inserting molecules exhibit fast or slow flip-flop. Fast flip-flop induces a 9 10 11235 large, visible increase in membrane area and fluctuations [75, 77, 78]. Slow flip-flop molecules 12 13236 induce an increase in the membrane spontaneous curvature due to their asymmetric incorporation into 14 15237 the outer leaflet of the membrane, resulting in tubulation [79-81], increased membrane tension and 16 17 18238 eventually rupture [75, 77]. Interestingly, the incorporation of simpler amphiphilic molecules or 19 20239 micelles leads to vesicle growth, often followed by division [70, 82], which is in line with predictions 21 22 23240 for phospholipids [83]. The limitation of detergent-like molecules is that at high concentrations, they 24 25241 lead to membrane permeabilization and eventual solubilization, depending on the membrane phase 26 27 28242 state and the CMC of the surfactant [84, 85]. In general, although this mixed approach provides a 29 30243 tangible system for studies of liposome growth, there is major conceptual drawback in the context of 31 32244 minimal cells – such microcompartments would lose their chemical identity over several generations 33 34 35245 and therefore could not sustain a continuous life cycle. The same drawback applies to another 36 37246 approach, in which blends of phospholipids and fatty acids or cationic surfactants were fused by 38 39 40247 attractive electrostatic interactions [86]. 41 42 43248 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 11

65 Directed Growth of Biomimetic 1 Microcompartments 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24249 25 26 27250 Figure 2. Some approaches for establishing growth of giant vesicles. (A) Growth via uptake of fatty 28251 acid molecules, which remain in dynamic exchange with the external and internal GUV solution and 29 30252 exhibit fast flip-flop across the membrane. (B) Growth via fusion mediated by the formation of 31253 transmembrane metal-ligand complex (left) or fusion protein mimetics (right). (C) Electrofusion 32254 relies on poration of two opposing membranes exposed to strong electric fields. The phase contrast 33255 snapshots on the right illustrate the fusion of two GUVs brought in contact and aligned by means of 34256 AC field, followed by the application of a DC pulse; reproduced from [87]. Copyright (2006) 35 36257 National Academy of Sciences. The white arrow indicates the direction of the field. (D) Growth of 37258 giant vesicles can be established by efficient fusion of small liposomes with the GUV, where the two 38259 membrane types are oppositely charged. Fusion is illustrated with the cartoon (left) and the confocal 39260 images (right) of one negatively charged GUV (green) exposed to a solution of positively charged 40 41261 small liposomes (red). Upon fusion, Förster resonance energy transfer leads to decrease in the red 42262 signal and the GUV area increases significantly [88]. The change occurs within seconds. Reprinted 43263 from [88] Copyright (2019), with permission from Elsevier. Scale bars in (C, D) represent 20 µm. 44 45264 46 47 48 49265 4 Growth of lipid vesicles – a difficult task, solved by fusion 50 51 52266 The system of highest interest for mimicking the proliferation of living cells is the phospholipid 53 54 55267 vesicle, which grows as a result of the incorporation of phospholipids in its membrane and eventually 56 57268 divides. The importance of liposomes arises from the chemical analogy with modern cell membranes. 58 59 60269 The cellular self-reproduction found in nature involves very complex biochemical interactions and 61 62 63 64 12

65 Directed Growth of Biomimetic 1 Microcompartments 2 3270 machineries. According to the reductionist concept, research in the context of protobiology and 4 5 6271 bottom-up synthetic biology aims to establish this phenomenon in a minimal system, based on 7 8272 simpler physicochemical mechanisms in a proof-of-concept for the origin and understanding of life. 9 10 11273 So far, however, growth of pure lipid vesicles has proved to be a challenge. Due to the difficulty of 12 13274 experimental realization of the membrane formation subsystem, however, theoretical studies have 14 15275 preceded practical examples in the case of phospholipids. Different kinetic [89] and thermodynamic 16 17 18276 [83, 90, 91] aspects have been considered and summarized together with other, chemoton-like 19 20277 approaches in a recent review [92]. On the practical side, the self-reproduction of a biomembrane, i.e. 21 22 23278 a phospholipid vesicle building its own membrane, has been addressed several times, but with only 24 25279 modest success so far. 26 27 28 29280 4.1 Growth via uptake of synthesized membrane components 30 31 32281 Efforts for synthesizing membrane components in liposomes date back to the early 90s. In particular, 33 34 35282 Luisi and coworkers assembled a four-enzyme cascade for phosphatidylcholine (PC) synthesis and 36 37283 deduced its incorporation into the membrane based on geometrical considerations (eventually the 38 39284 liposome size decreased, which was ascribed to the higher spontaneous curvature triggered by the 40 41 42285 newly synthesized short-chain PC – a partially undesired outcome with respect to growth) [93]. The 43 44286 authors did not unequivocally determine, whether the enzymes were present in the interior, whereby 45 46 47287 significant dimensional limitations could arise from the small liposome radius (23–26 nm), but they 48 49288 ascertained the enzyme association with the membrane to maintain an active form. In another study, 50 51 52289 the synthesis of phosphatidic acid by acyltransferases, expressed via cell-free methods in vesicles, 53 54290 was demonstrated but no growth was observed [94]. The latter observation was ascribed to the low 55 56291 amount of product due to the limited encapsulation, which will be difficult to overcome unless 57 58 59292 efficient transport mechanisms for the various precursors are established. Recently, this cell-free 60 61293 approach was extended to a more comprehensive pathway, starting from acyl-CoA and glycerol-3- 62 63 64 13

65 Directed Growth of Biomimetic 1 Microcompartments 2 3294 phosphate, and the liposome compartments were equipped with α-hemolysin to facilitate uptake of 4 5 6295 small polar molecules [95]. Although the authors did not specifically follow the vesicle growth, they 7 8296 identified practical barriers causing the low lipid synthesis rate, speculated about transport 9 10 11297 mechanisms for acyl-CoA and discussed relevant crowding and confinement effects. Another 12 13298 example in a similar direction was the biochemical synthesis of palmitate based on a FAS type I 14 15299 enzyme, encapsulated in liposomes [96], where vesicle growth has indeed been observed but the 16 17 18300 incorporation of foreign membrane components remains a conceptual problem. 19 20 21301 To circumvent protein synthesis issues, inherent to cell-free systems, an eight-enzyme cascade for the 22 23 24302 biosynthesis of phospholipids from fatty acids and glycerol 3-phosphate as building blocks has 25 26303 recently been assembled in vitro [97]. In the latter study, the enzymes and precursors were not 27 28 29304 encapsulated and synthesis occurred outside the liposomes. Experiments measuring the dequenching 30 31305 of rhodamine dye showed an approximately 30% membrane expansion, which was subsequently 32 33 34306 limited by the depletion of substrate. The growth was ascribed to the incorporation of oleic acid into 35 36307 the preformed liposomes, while the subsequent biosynthetic conversion to phosphatidic acid did not 37 38308 contribute to further expansion. Despite these limitations, the reconstitution of the biochemical 39 40 41309 membrane formation machinery is undisputedly a landmark in the area of bottom-up synthetic 42 43310 biology. 44 45 46 47311 The continuous search for more realistic growing protocell models has led to the use of phospholipid- 48 49312 like molecules, supplied from outside, instead of internally synthesized. The Sugawara group 50 51 52313 demonstrated growth, division and amplified DNA distribution in daughter vesicles [98]. This was 53 54314 achieved by adding a cationic precursor, which was hydrolyzed into the membrane lipid by an 55 56315 embedded catalyst. Notably, DNA amplification enhanced the efficiency of both growth and 57 58 59316 division. However, the elegant process was ultimately exhausted because it was limited by the 60 61317 consumption of DNA precursors inside the vesicles and the dilution of membrane catalyst and 62 63 64 14

65 Directed Growth of Biomimetic 1 Microcompartments 2 3318 phospholipids. Similar approaches were undertaken by the replacement of the complex biochemical 4 5 6319 lipid synthesis pathway with a self-reproducing autocatalyst, which resulted in triazole phospholipids 7 8320 formation and membrane growth [99]. Both cases seem to function with aggregates and multilamellar 9 10 11321 vesicles, and the exotic chemistry of the membranes might hinder interactions with some biological 12 13322 species (e.g. sensitive membrane proteins), thus potentially limiting other applications in the context 14 15323 of minimal cells. 16 17 18 19324 4.2 Growth via vesicle fusion 20 21 22325 In the case of authentic phospholipids, the only established and practical approach for achieving 23 24 25326 growth until now seems to be via vesicle fusion. Membrane fusion is a ubiquitous process in biology 26 27327 and it is fundamental for a number of cellular processes. It involves merging of two otherwise 28 29 30328 separated membranes, forming a compartment, whose area is the sum of the two fusing bilayers. In 31 32329 other words, increase in membrane area inevitably accompanies fusion, although the former is 33 34 35330 scarcely reported because the vast majority of fusion assays are insensitive to it. Driving forces of 36 37331 various nature have been employed to trigger fusion, some approaches are illustrated in Figure 2. 38 39332 These range from simple physicochemical triggers such as electrostatic interactions [100-102] and 40 41 42333 volume depletion [103] to biochemical approaches, relying on natural protein fusion machinery 43 44334 [104], as well as combinations thereof [105]. Furthermore, fusion methods have been expanded to 45 46 47335 biomimetic strategies [87, 106] in addition to more exotic approaches, involving the use of light (to 48 49336 heat nanoparticles [107] or isomerize azocompounds [108]) or electric fields [87, 109, 110] to perturb 50 51 52337 and porate the bilayer. In some of the examples fusion has led to spontaneous budding, correlating to 53 54338 predictions made by theoretical models [90], which has been put in the context of growth and 55 56339 division cycles for self-reproduction [109]. Presumably, the simplest mechanism to induce membrane 57 58 59340 fusion is based on membrane tension as has been elucidated by molecular simulations [111]. 60 61 62 63 64 15

65 Directed Growth of Biomimetic 1 Microcompartments 2 3341 Membrane fusion of small liposomes has been used for decades [112, 113] and only more recently 4 5 6342 fusion using giant vesicles has been reported [114]. In cells, membrane fusion is mediated by specific 7 8343 interactions between fusion proteins and such fusion reactions have been reconstituted in synthetic 9 10 11344 systems, in which small liposomes fuse with giant vesicles [104, 115, 116]. However, possibly due to 12 13345 the low fusion efficiency of protein-reconstituted systems or the low protein density, vesicle growth 14 15346 has not been detected. Fusion mediated by pH-sensitive lipids has also resulted in undetectable area 16 17 18347 increase [117]. In contrast, when mediated by high charge density, the fusion of positively-charged 19 20348 liposomes induces extensive area increase of the negative GUVs [88, 118], see also Figure 2D. 21 22 23349 Although very efficient, the process is limited by eventual charge neutralization. 24 25 26350 The fusion of two GUVs can be induced and observed also in a more controllable way. This can be 27 28 29351 achieved by bringing a pair of vesicles in contact by manipulation with ultramicroelectrodes [119], 30 31352 by trapping them within microfluidic devices [120], by applying an electric field [87, 110], or by 32 33 34353 direct manipulation using micropipettes [87, 121] or optical tweezers [107, 122]. After initial contact, 35 36354 fusion can be triggered by the application of a strong DC pulse [87, 123], via ligand-mediated ion 37 38355 interactions [87], transmembrane domains of fusogenic proteins [124] or through localized 39 40 41356 nanoparticle-mediated [107] or laser-mediated [122] heating. The common feature of all these studies 42 43357 is that an increase in membrane area is clearly observable. Although GUV fusion may result in 44 45 46358 visible area increase, the resulting area has never been quantified and fusion efficiency has been 47 48359 determined by changes in the measured fluorescence rather than the measured volume changes of the 49 50 51360 vesicle to characterize vesicular growth. Generally, the excess area results in increased membrane 52 53361 fluctuations and formation of membrane folds as a consequence of volume expansion lagging behind 54 55362 the area growth. The GUV-GUV fusion approach in its variations has been used to mix chemical 56 57 58363 precursors for nanoparticle formation inside the vesicular compartment [119, 122, 123], to 59 60 61 62 63 64 16

65 Directed Growth of Biomimetic 1 Microcompartments 2 3364 controllably obtain vesicles with well-defined compositions [121], and to unravel the formation of 4 5 6365 the hemifusion diaphragm [124, 125] and the rapid nature of the fusion neck expansion [87]. 7 8 9366 Importantly, fusion is not limited to pure phospholipid GUVs, but can be also extended to fusion of 10 11 12367 GUVs and cells [126, 127] or GUVs made of synthetic polymers. In contrast to liposomes, fusion of 13 14368 polymersomes is limited by the restricted mobility and high stability of the polymer membranes 15 16369 [128-130] associated with slowed down polymer rearrangement, which is a prerequisite for processes 17 18 19370 such as fusion. Some strategies have been successful in inducing fusion events in polymer GUVs, 20 21371 including exposure of vesicles to osmotic pressure gradients [128], ultrasound treatment [131], 22 23 24372 membrane perturbation by azobenzene photoisomerization [132] and addition of membrane 25 26373 disruptive agents into the dispersing medium [133]. 27 28 29 30374 In relation to bottom-up synthetic biology and protobiology, although fusion has been traditionally 31 32375 considered from a biophysical and biochemical perspective, spontaneous liposome growth could in 33 34 35376 principle occur (or could have occurred in the evolutionary history) by the discrete steps of vesicle 36 37377 fusion [134]. In addition, apart from a mechanism for increasing the membrane area, it has been 38 39378 speculated as a scenario, in which two systems combine their properties and thus increase the degree 40 41 42379 of biocomplexity [135], for instance considering that vesicular trafficking is a major transport 43 44380 mechanism for proteins in natural living cells [136]. This aspect was used to supply feeding 45 46 47381 components and trigger expression in the case of fusion of oppositely charged vesicles [137]. 48 49 50382 5 Synthetic chemistry enables new potential approaches for growth 51 52 53 54383 In parallel to the continuing efforts to form and grow new types of membraneless 55 56384 microcompartments, enable efficient fusion via new approaches, and ideally reconstitute the natural 57 58 59385 machinery for phospholipid synthesis, there is another possible pathway, which brings synthetic 60 61386 chemistry in the foreground of bottom-up synthetic biology. This aspect has been already partially 62 63 64 17

65 Directed Growth of Biomimetic 1 Microcompartments 2 3387 addressed by the demonstration of life-like properties of dichloromethane and nitrobenzene oil 4 5 6388 droplets, which grow, divide and partition cargo [138, 139]. The versatility of synthetic reactions was 7 8389 also used to augment phospholipid vesicles and make them grow [98, 99] as noted above. With 9 10 11390 respect to this, advances in the preparation of functional amphiphilic block copolymers provide 12 13391 powerful synthetic tools for the engineering of artificial compartments that could grow upon 14 15392 stimulation by light, temperature or pH. Here the stimuli would induce the build-up of membrane 16 17 18393 stresses that could relax through fusion events. In this way the chemistry of living systems could be 19 20394 extended far beyond natural building blocks, enabling not only the reverse engineering or re- 21 22 23395 engineering, but also the de novo engineering of life. Besides a powerful toolbox, which may enable 24 25396 futuristic applications, the latter concept of synthetic (in its conventional semantics of chemical) has 26 27 28397 resorting implications on the understanding of life as a display of self-organization and will expand 29 30398 our search criteria for other forms of life. 31 32 33 34399 6 Conflict of Interest 35 36400 The authors declare that the research was conducted in the absence of any commercial or financial 37401 relationships that could be construed as a potential conflict of interest. 38 39402 7 Author Contributions 40 41 42403 All authors listed have made a substantial, direct and intellectual contribution to the work, and 43404 approved it for publication. 44 45405 8 Funding 46 47406 This work is part of the MaxSynBio consortium, which is jointly funded by the Federal Ministry of 48 49407 Education and Research of Germany (BMBF) and the Max Planck Society (MPG). 50 51408 9 References 52 53409 [1] T. Ganti, Organization of Chemical-Reactions into Dividing and Metabolizing Units - 54410 Chemotons, Biosystems 1975, 7, 15-21. 55 56411 [2] T. Ganti, itself, Journal of Theoretical Biology 1997, 187, 583-593. 57412 [3] P.L. Luisi, Autopoiesis: a review and a reappraisal, Naturwissenschaften 2003, 90, 49-59. 58413 [4] F.G. Varela, H.R. Maturana, R. Uribe, Autopoiesis: the organization of living systems, its 59414 characterization and a model, Biosystems 1974, 5, 187-196. 60415 [5] A.I. Oparin, Origin of life, Macmillan, New York, 1938. 61 62 63 64 18

65 Directed Growth of Biomimetic 1 Microcompartments 2 3416 [6] A.I. Oparin, K.L. Gladilin, Evolution of self-assembly of probionts, Biosystems 1980, 12, 133- 4417 145. 5 6418 [7] S.F. Banani, H.O. Lee, A.A. Hyman, M.K. Rosen, Biomolecular condensates: organizers of 7419 cellular biochemistry, Nature Reviews Molecular Cell Biology 2017, 18, 285-298. 8420 [8] C.P. Brangwynne, C.R. Eckmann, D.S. Courson, A. Rybarska, C. Hoege, J. Gharakhani, F. 9421 Julicher, A.A. Hyman, Germline P Granules Are Liquid Droplets That Localize by Controlled 10 11422 Dissolution/Condensation, Science 2009, 324, 1729-1732. 12423 [9] C.P. Brangwynne, T.J. Mitchison, A.A. Hyman, Active liquid-like behavior of nucleoli 13424 determines their size and shape in Xenopus laevis oocytes, Proceedings of the National Academy of 14425 Sciences of the United States of America 2011, 108, 4334-4339. 15426 [10] A.A. Hyman, C.A. Weber, F. Juelicher, Liquid-Liquid Phase Separation in Biology, in: R. 16 17427 Schekman, R. Lehmann (Eds.) Annual Review of Cell and Developmental Biology, Vol 302014, pp. 18428 39-58. 19429 [11] P.L. Li, S. Banjade, H.C. Cheng, S. Kim, B. Chen, L. Guo, M. Llaguno, J.V. Hollingsworth, 20430 D.S. King, S.F. Banani, P.S. Russo, Q.X. Jiang, B.T. Nixon, M.K. Rosen, Phase transitions in the 21431 assembly of multivalent signalling proteins, Nature 2012, 483, 336-U129. 22 23432 [12] Y. Shin, C.P. Brangwynne, Liquid phase condensation in cell physiology and disease, Science 24433 2017, 357. 25434 [13] X.L. Su, J.A. Ditlev, E.F. Hui, W.M. Xing, S. Banjade, J. Okrut, D.S. King, J. Taunton, M.K. 26435 Rosen, R.D. Vale, Phase separation of signaling molecules promotes T cell receptor signal 27 28436 transduction, Science 2016, 352, 595-599. 29437 [14] S. Elbaum-Garfinkle, Y. Kim, K. Szczepaniak, C.C.H. Chen, C.R. Eckmann, S. Myong, C.P. 30438 Brangwynne, The disordered P granule protein LAF-1 drives phase separation into droplets with 31439 tunable viscosity and dynamics, Proceedings of the National Academy of Sciences of the United 32440 States of America 2015, 112, 7189-7194. 33 34441 [15] S. Saha, C.A. Weber, M. Nousch, O. Adame-Arana, C. Hoege, M.Y. Hein, E. Osborne- 35442 Nishimura, J. Mahamid, M. Jahnel, L. Jawerth, A. Pozniakovski, C.R. Eckmann, F. Julicher, A.A. 36443 Hyman, Polar Positioning of Phase-Separated Liquid Compartments in Cells Regulated by an mRNA 37444 Competition Mechanism, Cell 2016, 166, 1572-+. 38 39445 [16] K.A. Burke, A.M. Janke, C.L. Rhine, N.L. Fawzi, Residue-by-Residue View of In Vitro FUS 40446 Granules that Bind the C-Terminal Domain of RNA Polymerase II, Molecular Cell 2015, 60, 231- 41447 241. 42448 [17] T. Murakami, S. Qamar, J.Q. Lin, G.S.K. Schierle, E. Rees, A. Miyashita, A.R. Costa, R.B. 43449 Dodd, F.T.S. Chan, C.H. Michel, D. Kronenberg-Versteeg, Y. Li, S.P. Yang, Y. Wakutani, W. 44 45450 Meadows, R.R. Ferry, L. Dong, G.G. Tartaglia, G. Favrin, W.L. Lin, D.W. Dickson, M. Zhen, D. 46451 Ron, G. Schmitt-Ulms, P.E. Fraser, N.A. Shneider, C. Holt, M. Vendruscolo, C.F. Kaminski, P. St 47452 George-Hyslop, ALS/FTD Mutation-Induced Phase Transition of FUS Liquid Droplets and 48453 Reversible Hydrogels into Irreversible Hydrogels Impairs RNP Granule Function, Neuron 2015, 88, 49 50454 678-690. 51455 [18] A. Patel, H.O. Lee, L. Jawerth, S. Maharana, M. Jahnel, M.Y. Hein, S. Stoynov, J. Mahamid, S. 52456 Saha, T.M. Franzmann, A. Pozniakovski, I. Poser, N. Maghelli, L.A. Royer, M. Weigert, E.W. 53457 Myers, S. Grill, D. Drechsel, A.A. Hyman, S. Alberti, A Liquid-to-Solid Phase Transition of the ALS 54458 Protein FUS Accelerated by Disease Mutation, Cell 2015, 162, 1066-1077. 55 56459 [19] J.B. Woodruff, B.F. Gomes, P.O. Widlund, J. Mahamid, A. Honigmann, A.A. Hyman, The 57460 Centrosome Is a Selective Condensate that Nucleates Microtubules by Concentrating Tubulin, Cell 58461 2017, 169, 1066-+. 59 60 61 62 63 64 19

65 Directed Growth of Biomimetic 1 Microcompartments 2 3462 [20] M. Feric, N. Vaidya, T.S. Harmon, D.M. Mitrea, L. Zhu, T.M. Richardson, R.W. Kriwacki, R.V. 4463 Pappu, C.P. Brangwynne, Coexisting Liquid Phases Underlie Nucleolar Subcompartments, Cell 5 6464 2016, 165, 1686-1697. 7465 [21] A. Hubstenberger, S.L. Noble, C. Cameron, T.C. Evans, Repressors, an RNA 8466 Helicase, and Developmental Cues Control RNP Phase Transitions during Early Development, 9467 Developmental Cell 2013, 27, 161-173. 10 11468 [22] T.M. Franzmann, M. Jahnel, A. Pozniakovsky, J. Mahamid, A.S. Holehouse, E. Nuske, D. 12469 Richter, W. Baumeister, S.W. Grill, R.V. Pappu, A.A. Hyman, S. Alberti, Phase separation of a yeast 13470 prion protein promotes cellular fitness, Science 2018, 359, 47-+. 14471 [23] J. Wang, J.M. Choi, A.S. Holehouse, H.O. Lee, X.J. Zhang, M. Jahnel, S. Maharana, R. 15472 Lemaitre, A. Pozniakovsky, D. Drechsel, I. Poser, R.V. Pappu, S. Alberti, A.A. Hyman, A Molecular 16 17473 Grammar Governing the Driving Forces for Phase Separation of Prion-like RNA Binding Proteins, 18474 Cell 2018, 174, 688-+. 19475 [24] S. Boeynaems, S. Alberti, N.L. Fawzi, T. Mittag, M. Polymenidou, F. Rousseau, J. 20476 Schymkowitz, J. Shorter, B. Wolozin, L. Van den Bosch, P. Tompa, M. Fuxreiter, Protein Phase 21477 Separation: A New Phase in Cell Biology, Trends in Cell Biology 2018, 28, 420-435. 22 23478 [25] T.J. Nott, E. Petsalaki, P. Farber, D. Jervis, E. Fussner, A. Plochowietz, T.D. Craggs, D.P. 24479 Bazett-Jones, T. Pawson, J.D. Forman-Kay, A.J. Baldwin, Phase Transition of a Disordered Nuage 25480 Protein Generates Environmentally Responsive Membraneless Organelles, Molecular Cell 2015, 57, 26481 936-947. 27 28482 [26] T.S. Harmon, A.S. Holehouse, M.K. Rosen, R.V. Pappu, Intrinsically disordered linkers 29483 determine the interplay between phase separation and gelation in multivalent proteins, Elife 2017, 6. 30484 [27] A. Molliex, J. Temirov, J. Lee, M. Coughlin, A.P. Kanagaraj, H.J. Kim, T. Mittag, J.P. Taylor, 31485 Phase Separation by Low Complexity Domains Promotes Stress Granule Assembly and Drives 32486 Pathological Fibrillization, Cell 2015, 163, 123-133. 33 34487 [28] D. Zwicker, A.A. Hyman, F. Julicher, Suppression of Ostwald ripening in active emulsions, 35488 Physical Review E 2015, 92. 36489 [29] D. Zwicker, R. Seyboldt, C.A. Weber, A.A. Hyman, F. Julicher, Growth and division of active 37490 droplets provides a model for protocells, Nature Physics 2017, 13, 408-413. 38 39491 [30] T. Franzmann, S. Alberti, Prion-like low-complexity sequences: Key regulators of protein 40492 solubility and phase behavior, J Biol Chem 2018. 41493 [31] K.M. Ruff, S. Roberts, A. Chilkoti, R.V. Pappu, Advances in Understanding Stimulus 42494 Responsive Phase Behavior of Intrinsically Disordered Protein Polymers, J Mol Biol 2018. 43495 [32] R. Chollakup, W. Smitthipong, C.D. Eisenbach, M. Tirrell, Phase Behavior and Coacervation of 44 45496 Aqueous Poly(acrylic acid)-Poly(allylamine) Solutions, Macromolecules 2010, 43, 2518-2528. 46497 [33] T.Y.D. Tang, C.R.C. Hak, A.J. Thompson, M.K. Kuimova, D.S. Williams, A.W. Perriman, S. 47498 Mann, Fatty acid membrane assembly on coacervate microdroplets as a step towards a hybrid 48499 protocell model, Nature Chemistry 2014, 6, 527-533. 49 50500 [34] K.A. Black, D. Priftis, S.L. Perry, J. Yip, W.Y. Byun, M. Tirrell, Protein Encapsulation via 51501 Polypeptide Complex Coacervation, Acs Macro Letters 2014, 3, 1088-1091. 52502 [35] D. Priftis, K. Megley, N. Laugel, M. Tirrell, Complex coacervation of poly(ethylene- 53503 imine)/polypeptide aqueous solutions: Thermodynamic and rheological characterization, Journal of 54504 and Interface Science 2013, 398, 39-50. 55 56505 [36] D. Priftis, M. Tirrell, Phase behaviour and complex coacervation of aqueous polypeptide 57506 solutions, Soft Matter 2012, 8, 9396-9405. 58507 [37] W.M. Aumiller, C.D. Keating, Phosphorylation-mediated RNA/peptide complex coacervation as 59508 a model for intracellular liquid organelles, Nature Chemistry 2016, 8, 129-137. 60 61 62 63 64 20

65 Directed Growth of Biomimetic 1 Microcompartments 2 3509 [38] J.R. Vieregg, M. Lueckheide, A.B. Marciel, L. Leon, A.J. Bologna, J.R. Rivera, M.V. Tirrell, 4510 Oligonucleotide-Peptide Complexes: Phase Control by Hybridization, Journal of the American 5 6511 Chemical Society 2018, 140, 1632-1638. 7512 [39] S. Koga, D.S. Williams, A.W. Perriman, S. Mann, Peptide-nucleotide microdroplets as a step 8513 towards a membrane-free protocell model, Nature Chemistry 2011, 3, 720-724. 9514 [40] R. Dimova, R. Lipowsky, Lipid membranes in contact with aqueous phases of polymer 10 11515 solutions, Soft Matter 2012, 8, 6409-6415. 12516 [41] C.D. Keating, Aqueous Phase Separation as a Possible Route to Compartmentalization of 13517 Biological Molecules, Accounts of Chemical Research 2012, 45, 2114-2124. 14518 [42] C.W. Pak, M. Kosno, A.S. Holehouse, S.B. Padrick, A. Mittal, R. Ali, A.A. Yunus, D.R. Liu, 15519 R.V. Pappu, M.K. Rosen, Sequence Determinants of Intracellular Phase Separation by Complex 16 17520 Coacervation of a Disordered Protein, Molecular Cell 2016, 63, 72-85. 18521 [43] V.N. Uversky, I.M. Kuznetsova, K.K. Turoverov, B. Zaslavsky, Intrinsically disordered proteins 19522 as crucial constituents of cellular aqueous two phase systems and coacervates, Febs Letters 2015, 20523 589, 15-22. 21524 [44] E. Spruijt, J. Sprakel, M.A.C. Stuart, J. van der Gucht, Interfacial tension between a complex 22 23525 coacervate phase and its coexisting aqueous phase, Soft Matter 2010, 6, 172-178. 24526 [45] Y.G. Liu, R. Lipowsky, R. Dimova, Concentration Dependence of the Interfacial Tension for 25527 Aqueous Two-Phase Polymer Solutions of Dextran and Polyethylene Glycol, Langmuir 2012, 28, 26528 3831-3839. 27 28529 [46] D.S. Williams, S. Koga, C.R.C. Hak, A. Majrekar, A.J. Patil, A.W. Perriman, S. Mann, 29530 Polymer/nucleotide droplets as bio-inspired functional micro-compartments, Soft Matter 2012, 8, 30531 6004-6014. 31532 [47] E.A. Frankel, P.C. Bevilacqua, C.D. Keating, Polyamine/Nucleotide Coacervates Provide Strong 32533 Compartmentalization of Mg2+, Nucleotides, and RNA, Langmuir 2016, 32, 2041-2049. 33 34534 [48] T.Y.D. Tang, M. Antognozzi, J.A. Vicary, A.W. Perriman, S. Mann, Small-molecule uptake in 35535 membrane-free peptide/nucleotide protocells, Soft Matter 2013, 9, 7647-7656. 36536 [49] M.S. Long, C.D. Jones, M.R. Helfrich, L.K. Mangeney-Slavin, C.D. Keating, Dynamic 37537 microcompartmentation in synthetic cells, Proc. Natl. Acad. Sci. U. S. A. 2005, 102, 5920-5925. 38 39538 [50] E. Sokolova, E. Spruijt, M.M.K. Hansen, E. Dubuc, J. Groen, V. Chokkalingam, A. Piruska, 40539 H.A. Heus, W.T.S. Huck, Enhanced rates in membrane-free protocells formed by 41540 coacervation of cell lysate, Proceedings of the National Academy of Sciences of the United States of 42541 America 2013, 110, 11692-11697. 43542 [51] T.Y.D. Tang, D. van Swaay, A. deMello, J.L.R. Anderson, S. Mann, In vitro 44 45543 within membrane-free coacervate protocells, Chemical Communications 2015, 51, 11429-11432. 46544 [52] K.K. Nakashima, J.F. Baaij, E. Spruijt, Reversible generation of coacervate droplets in an 47545 enzymatic network, Soft Matter 2018, 14, 361-367. 48546 [53] P. Stano, P.L. Luisi, Achievements and open questions in the self-reproduction of vesicles and 49 50547 synthetic minimal cells, Chem Commun (Camb) 2010, 46, 3639-3653. 51548 [54] M.M. Hanczyc, S.M. Fujikawa, J.W. Szostak, Experimental models of primitive cellular 52549 compartments: Encapsulation, growth, and division, Science 2003, 302, 618-622. 53550 [55] D. Deamer, J.P. Dworkin, S.A. Sandford, M.P. Bernstein, L.J. Allamandola, The first cell 54551 membranes, 2002, 2, 371-381. 55 56552 [56] N.R.C. Steering Group for the Workshop on Size Limits of Very Small Microorganisms, Size 57553 Limits of Very Small Microorganisms: Proceedings of a Workshop, National Academies Press (US) 58 59554 Copyright 1999 by the National Academy of Sciences. All rights reserved., Washington (DC), 1999. 60 61 62 63 64 21

65 Directed Growth of Biomimetic 1 Microcompartments 2 3555 [57] T.P. de Souza, P. Stano, P.L. Luisi, The Minimal Size of Liposome-Based Model Cells Brings 4556 about a Remarkably Enhanced Entrapment and Protein Synthesis, Chembiochem 2009, 10, 1056- 5 6557 1063. 7558 [58] R. Dimova, S. Aranda, N. Bezlyepkina, V. Nikolov, K.A. Riske, R. Lipowsky, A practical guide 8559 to giant vesicles. Probing the membrane nanoregime via optical microscopy, J. Phys.: Condens. 9560 Matter 2006, 18, S1151-S1176. 10 11561 [59] R. Dimova, Giant Vesicles: A Biomimetic Tool for Membrane Characterization, in: A. Iglič 12562 (Ed.) Advances in Planar Lipid Bilayers and Liposomes, Academic Press2012, pp. 1-50. 13563 [60] P. Walde, K. Cosentino, H. Engel, P. Stano, Giant Vesicles: Preparations and Applications, 14564 ChemBioChem 2010, 11, 848-865. 15565 [61] S.F. Fenz, K. Sengupta, Giant vesicles as cell models, Integr Biol-Uk 2012, 4, 982-995. 16 17566 [62] K. Morigaki, P. Walde, Fatty acid vesicles, Current Opinion in Colloid & Interface Science 18567 2007, 12, 75-80. 19568 [63] I.A. Chen, R.W. Roberts, J.W. Szostak, The Emergence of Competition Between Model 20569 Protocells, Science 2004, 305, 1474. 21570 [64] P. Walde, R. Wick, M. Fresta, A. Mangone, P.L. Luisi, Autopoietic Self-Reproduction of Fatty 22 23571 Acid Vesicles, J. Am. Chem. Soc. 1994, 116, 11649-11654. 24572 [65] E. Blochliger, M. Blocher, P. Walde, P.L. Luisi, Matrix effect in the size distribution of fatty 25573 acid vesicles, Journal of Physical Chemistry B 1998, 102, 10383-10390. 26574 [66] I.A. Chen, J.W. Szostak, A kinetic study of the growth of fatty acid vesicles, Biophys J 2004, 87, 27 28575 988-998. 29576 [67] A.J. Markvoort, N. Pfleger, R. Staffhorst, P.A. Hilbers, R.A. van Santen, J.A. Killian, B. de 30577 Kruijff, Self-reproduction of fatty acid vesicles: a combined experimental and simulation study, 31578 Biophys J 2010, 99, 1520-1528. 32579 [68] C. Hentrich, J.W. Szostak, Controlled growth of filamentous fatty acid vesicles under flow, 33 34580 Langmuir 2014, 30, 14916-14925. 35581 [69] A.E. Engelhart, K.P. Adamala, J.W. Szostak, A simple physical mechanism enables homeostasis 36582 in primitive cells, Nature Chemistry 2016, 8, 448-453. 37583 [70] T.F. Zhu, J.W. Szostak, Coupled Growth and Division of Model Protocell Membranes, Journal 38 39584 of the American Chemical Society 2009, 131, 5705-5713. 40585 [71] A. Shioi, T.A. Hatton, Model for formation and growth of vesicles in mixed anionic/cationic 41586 (SOS/CTAB) surfactant systems, Langmuir 2002, 18, 7341-7348. 42587 [72] N. Berclaz, M. Muller, P. Walde, P.L. Luisi, Growth and transformation of vesicles studied by 43588 ferritin labeling and cryotransmission electron microscopy, Journal of Physical Chemistry B 2001, 44 45589 105, 1056-1064. 46590 [73] T.P. de Souza, M. Holzer, P. Stano, F. Steiniger, S. May, R. Schubert, A. Fahr, P.L. Luisi, New 47591 Insights into the Growth and Transformation of Vesicles: A Free-Flow Electrophoresis Study, 48592 Journal of Physical Chemistry B 2015, 119, 12212-12223. 49 50593 [74] M.L. Rogerson, B.H. Robinson, S. Bucak, P. Walde, Kinetic studies of the interaction of fatty 51594 acids with phosphatidylcholine vesicles (liposomes), Colloids Surf B Biointerfaces 2006, 48, 24-34. 52595 [75] P. Peterlin, V. Arrigler, K. Kogej, S. Svetina, P. Walde, Growth and shape transformations of 53596 giant phospholipid vesicles upon interaction with an aqueous oleic acid suspension, Chemistry and 54597 Physics of Lipids 2009, 159, 67-76. 55 56598 [76] J. Dervaux, V. Noireaux, A.J. Libchaber, Growth and instability of a phospholipid vesicle in a 57599 bath of fatty acids, The European Physical Journal Plus 2017, 132. 58600 [77] T.P. Sudbrack, N.L. Archilha, R. Itri, K.A. Riske, Observing the Solubilization of Lipid Bilayers 59601 by Detergents with Optical Microscopy of GUVs, Journal of Physical Chemistry B 2011, 115, 269- 60602 277. 61 62 63 64 22

65 Directed Growth of Biomimetic 1 Microcompartments 2 3603 [78] V.N. Georgiev, A. Grafmüller, D. Bléger, S. Hecht, S. Kunstmann, S. Barbirz, R. Lipowsky, R. 4604 Dimova, Area Increase and Budding in Giant Vesicles Triggered by Light: Behind the Scene, 5 6605 Advanced Science 2018, 5, 1800432. 7606 [79] R. Dasgupta, M.S. Miettinen, N. Fricke, R. Lipowsky, R. Dimova, The glycolipid GM1 reshapes 8607 asymmetric biomembranes and giant vesicles by curvature generation, Proc. Natl. Acad. Sci. U. S. A. 9608 2018, 115, 5756–5761. 10 11609 [80] R. Lipowsky, Spontaneous tubulation of membranes and vesicles reveals membrane tension 12610 generated by spontaneous curvature, Faraday Discuss. 2013, 161, 305-331. 13611 [81] P. Bassereau, R. Jin, T. Baumgart, M. Deserno, R. Dimova, V.A. Frolov, P.V. Bashkirov, H. 14612 Grubmüller, R. Jahn, H.J. Risselada, L. Johannes, M.M. Kozlov, R. Lipowsky, T.J. Pucadyil, W.F. 15613 Zeno, J.C. Stachowiak, D. Stamou, A. Breuer, L. Lauritsen, C. Simon, C. Sykes, G.A. Voth, T.R. 16 17614 Weikl, The 2018 biomembrane curvature and remodeling roadmap, J. Phys. D: Appl. Phys. 2018, 51, 18615 343001. 19616 [82] Y. Inaoka, M. Yamazaki, Vesicle fission of giant unilamellar vesicles of liquid-ordered-phase 20617 membranes induced by amphiphiles with a single long hydrocarbon chain, Langmuir 2007, 23, 720- 21618 728. 22 23619 [83] B. Bozic, S. Svetina, A relationship between membrane properties forms the basis of a 24620 selectivity mechanism for vesicle self-reproduction, European Biophysics Journal with Biophysics 25621 Letters 2004, 33, 565-571. 26622 [84] D. Lichtenberg, H. Ahyayauch, F.M. Goni, The Mechanism of Detergent Solubilization of Lipid 27 28623 Bilayers, Biophysical Journal 2013, 105, 289-299. 29624 [85] B. Mattei, R.B. Lira, K.R. Perez, K.A. Riske, Membrane permeabilization induced by Triton X- 30625 100: The role of membrane phase state and edge tension, Chemistry and Physics of Lipids 2017, 202, 31626 28-37. 32627 [86] F. Caschera, P. Stano, P.L. Luisi, Reactivity and fusion between cationic vesicles and fatty acid 33 34628 anionic vesicles, Journal of Colloid and Interface Science 2010, 345, 561-565. 35629 [87] C.K. Haluska, K.A. Riske, V. Marchi-Artzner, J.M. Lehn, R. Lipowsky, R. Dimova, Time scales 36630 of membrane fusion revealed by direct imaging of vesicle fusion with high temporal resolution, Proc 37631 Natl Acad Sci U S A 2006, 103, 15841-15846. 38 39632 [88] R.B. Lira, T. Robinson, R. Dimova, K.A. Riske, Charge-dependent membrane fusion in a pure 40633 lipid system: a single giant vesicle study, Biophys J 2018, (in press). 41634 [89] V.P. Zhdanov, B. Kasemo, Lipid-diffusion-limited kinetics of vesicle growth, Langmuir 2000, 42635 16, 7352-7354. 43636 [90] B. Bozic, S. Svetina, Vesicle self-reproduction: The involvement of membrane hydraulic and 44 45637 solute permeabilities, European Physical Journal E 2007, 24, 79-90. 46638 [91] D. Fanelli, A.J. McKane, Thermodynamics of vesicle growth and instability, Physical Review E 47639 2008, 78. 48640 [92] F. Mavelli, E. Altamura, L. Cassidei, P. Stano, Recent Theoretical Approaches to Minimal 49 50641 Artificial Cells, Entropy 2014, 16, 2488-2511. 51642 [93] P.K. Schmidli, P. Schurtenberger, P.L. Luisi, Liposome-mediated enzymatic-synthesis of 52643 phosphatidylcholine as an approach to self-replicating liposomes, Journal of the American Chemical 53644 Society 1991, 113, 8127-8130. 54645 [94] Y. Kuruma, P. Stano, T. Ueda, P.L. Luisi, A synthetic biology approach to the construction of 55 56646 membrane proteins in semi-synthetic minimal cells, Biochim Biophys Acta 2009, 1788, 567-574. 57647 [95] A. Scott, M.J. Noga, P. de Graaf, I. Westerlaken, E. Yildirim, C. Danelon, Cell-Free 58648 Phospholipid Biosynthesis by Gene-Encoded Enzymes Reconstituted in Liposomes, Plos One 2016, 59649 11. 60 61 62 63 64 23

65 Directed Growth of Biomimetic 1 Microcompartments 2 3650 [96] G. Murtas, Internal lipid synthesis and vesicle growth as a step toward self-reproduction of the 4651 minimal cell, Syst Synth Biol 2010, 4, 85-93. 5 6652 [97] M. Exterkate, A. Caforio, M.C.A. Stuart, A.J.M. Driessen, Growing Membranes In Vitro by 7653 Continuous Phospholipid Biosynthesis from Free Fatty Acids, Acs Synthetic Biology 2018, 7, 153- 8654 165. 9655 [98] K. Kurihara, M. Tamura, K. Shohda, T. Toyota, K. Suzuki, T. Sugawara, Self-reproduction of 10 11656 supramolecular giant vesicles combined with the amplification of encapsulated DNA, Nat Chem 12657 2011, 3, 775-781. 13658 [99] M.D. Hardy, J. Yang, J. Selimkhanov, C.M. Cole, L.S. Tsimring, N.K. Devaraj, Self- 14659 reproducing catalyst drives repeated phospholipid synthesis and membrane growth, Proceedings of 15660 the National Academy of Sciences of the United States of America 2015, 112, 8187-8192. 16 17661 [100] T. Tanaka, M. Yamazaki, Membrane fusion of giant unilamellar vesicles of neutral 18662 phospholipid membranes induced by La3+, Langmuir 2004, 20, 5160-5164. 19663 [101] D.P. Pantazatos, R.C. MacDonald, Directly observed membrane fusion between oppositely 20664 charged phospholipid bilayers, Journal of Membrane Biology 1999, 170, 27-38. 21665 [102] T. Sunami, F. Caschera, Y. Morita, T. Toyota, K. Nishimura, T. Matsuura, H. Suzuki, M.M. 22 23666 Hanczyc, T. Yomo, Detection of Association and Fusion of Giant Vesicles Using a Fluorescence- 24667 Activated Cell Sorter, Langmuir 2010, 26, 15098-15103. 25668 [103] B.R. Lentz, PEG as a tool to gain insight into membrane fusion, Eur Biophys J 2007, 36, 315- 26669 326. 27 28670 [104] D. Tareste, J. Shen, T.J. Melia, J.E. Rothman, SNAREpin/Munc18 promotes adhesion and 29671 fusion of large vesicles to giant membranes, Proceedings of the National Academy of Sciences of the 30672 United States of America 2008, 105, 2380-2385. 31673 [105] M. Kyoung, Y. Zhang, J. Diao, S. Chu, A.T. Brunger, Studying calcium-triggered vesicle 32674 fusion in a single vesicle-vesicle content and lipid-mixing system, Nat Protoc 2013, 8, 1-16. 33 34675 [106] Y.H.M. Chan, B. van Lengerich, S.G. Boxer, Lipid-anchored DNA mediates vesicle fusion as 35676 observed by lipid and content mixing, Biointerphases 2008, 3, FA17-FA21. 36677 [107] A. Rorvig-Lund, A. Bahadori, S. Semsey, P.M. Bendix, L.B. Oddershede, Vesicle Fusion 37678 Triggered by Optically Heated Gold Nanoparticles, Nano Letters 2015, 15, 4183-4188. 38 39679 [108] Y. Suzuki, K.H. Nagai, A. Zinchenko, T. Hamada, Photoinduced Fusion of 40680 Membranes, Langmuir 2017, 33, 2671-2676. 41681 [109] H. Terasawa, K. Nishimura, H. Suzuki, T. Matsuura, T. Yomo, Coupling of the fusion and 42682 budding of giant phospholipid vesicles containing macromolecules, Proc Natl Acad Sci U S A 2012, 43683 109, 5942-5947. 44 45684 [110] K.A. Riske, N. Bezlyepkina, R. Lipowsky, R. Dimova, Electrofusion of model lipid 46685 membranes viewed with high temporal resolution, Biophys. Rev. Lett. 2006, 1, 387-400. 47686 [111] J.C. Shillcock, R. Lipowsky, Tension-induced fusion of bilayer membranes and vesicles, Nat. 48687 Mater. 2005, 4, 225-228. 49 50688 [112] D.K. Struck, D. Hoekstra, R.E. Pagano, Use of resonance energy-transfer to monitor membrane 51689 fusion, Biochemistry 1981, 20, 4093-4099. 52690 [113] J. Wilschut, N. Duzgunes, R. Fraley, D. Papahadjopoulos, Studies on the mechanism of 53691 membrane-fusion - kinetics of calcium-ion induced fusion of phosphatidylserine vesicles followed by 54692 a new assay for mixing of aqueous vesicle contents, Biochemistry 1980, 19, 6011-6021. 55 56693 [114] G.H. Lei, R.C. MacDonald, Lipid bilayer vesicle fusion: Intermediates captured by high-speed 57694 microfluorescence spectroscopy, Biophysical Journal 2003, 85, 1585-1599. 58695 [115] A. Witkowska, R. Jahn, Rapid SNARE-Mediated Fusion of Liposomes and Chromaffin 59696 Granules with Giant Unilamellar Vesicles, Biophysical Journal 2017, 113, 1251-1259. 60 61 62 63 64 24

65 Directed Growth of Biomimetic 1 Microcompartments 2 3697 [116] T.-T. Kliesch, J. Dietz, L. Turco, P. Halder, E. Polo, M. Tarantola, R. Jahn, A. Janshoff, 4698 Membrane tension increases fusion efficiency of model membranes in the presence of SNAREs, Sci 5 6699 Rep 2017, 7, 12070. 7700 [117] S. Trier, J.R. Henriksen, T.L. Andresen, Membrane fusion of pH-sensitive liposomes - a 8701 quantitative study using giant unilamellar vesicles, Soft Matter 2011, 7, 9027-9034. 9702 [118] J. Solon, J. Pecreaux, P. Girard, M.C. Faure, J. Prost, P. Bassereau, Negative tension induced 10 11703 by lipid uptake, Phys. Rev. Lett. 2006, 97, 098103. 12704 [119] D.T. Chiu, C.F. Wilson, F. Ryttsen, A. Stromberg, C. Farre, A. Karlsson, S. Nordholm, A. 13705 Gaggar, B.P. Modi, A. Moscho, R.A. Garza-Lopez, O. Orwar, R.N. Zare, Chemical transformations 14706 in individual ultrasmall biomimetic containers, Science 1999, 283, 1892-1895. 15707 [120] T. Robinson, P.E. Verboket, K. Eyer, P.S. Dittrich, Controllable electrofusion of lipid vesicles: 16 17708 initiation and analysis of reactions within biomimetic containers, Lab on a Chip 2014, 14, 2852- 18709 2859. 19710 [121] N. Bezlyepkina, R.S. Gracia, P. Shchelokovskyy, R. Lipowsky, R. Dimova, Phase Diagram 20711 and Tie-Line Determination for the Ternary Mixture DOPC/eSM/Cholesterol, Biophysical Journal 21712 2013, 104, 1456-1464. 22 23713 [122] S. Kulin, R. Kishore, K. Helmerson, L. Locascio, Optical manipulation and fusion of liposomes 24714 as microreactors, Langmuir 2003, 19, 8206-8210. 25715 [123] P. Yang, R. Lipowsky, R. Dimova, Nanoparticle formation in giant vesicles: synthesis in 26716 biomimetic compartments, Small 2009, 5, 2033-2037. 27 28717 [124] J. Nikolaus, M. Stöckl, D. Langosch, R. Volkmer, A. Herrmann, Direct Visualization of Large 29718 and Protein-Free Hemifusion Diaphragms, Biophys. J. 2010, 98, 1192-1199. 30719 [125] J. Heuvingh, F. Pincet, S. Cribier, Hemifusion and fusion of giant vesicles induced by 31720 reduction of inter-membrane distance, The European Physical Journal E 2004, 14, 269-276. 32721 [126] A.C. Saito, T. Ogura, K. Fujiwara, S. Murata, S.M. Nomura, Introducing Micrometer-Sized 33 34722 Artificial Objects into Live Cells: A Method for Cell-Giant Unilamellar Vesicle Electrofusion, Plos 35723 One 2014, 9. 36724 [127] A. Strömberg, F. Ryttsén, D.T. Chiu, M. Davidson, P.S. Eriksson, C.F. Wilson, O. Orwar, R.N. 37725 Zare, Manipulating the genetic identity and biochemical surface properties of individual cells with 38 39726 electric-field-induced fusion, Proceedings of the National Academy of Sciences 2000, 97, 7. 40727 [128] B.M. Discher, Y.Y. Won, D.S. Ege, J.C.M. Lee, F.S. Bates, D.E. Discher, D.A. Hammer, 41728 Polymersomes: Tough vesicles made from diblock copolymers, Science 1999, 284, 1143-1146. 42729 [129] J.C.M. Lee, M. Santore, F.S. Bates, D.E. Discher, From membranes to melts, rouse to 43730 reptation: Diffusion in polymersome versus lipid bilayers, Macromolecules 2002, 35, 323-326. 44 45731 [130] R. Dimova, U. Seifert, B. Pouligny, S. Forster, H.G. Dobereiner, Hyperviscous diblock 46732 copolymer vesicles, Eur. Phys. J. E 2002, 7, 241-250. 47733 [131] Y.F. Zhou, D.Y. Yan, Real-time membrane fusion of giant polymer vesicles, Journal of the 48734 American Chemical Society 2005, 127, 10468-10469. 49 50735 [132] W. Su, Y.H. Luo, Q. Yan, S. Wu, K. Han, Q.J. Zhang, Y.Q. Gu, Y.M. Li, Photoinduced fusion 51736 of micro-vesicles self-assembled from azobenzene-containing amphiphilic diblock copolymers, 52737 Macromolecular Rapid Communications 2007, 28, 1251-1256. 53738 [133] T.P. Smart, C. Fernyhough, A.J. Ryan, G. Battaglia, Controlling Fusion and Aggregation in 54739 Polymersome Dispersions, Macromolecular Rapid Communications 2008, 29, 1855-1860. 55 56740 [134] J.W. Szostak, D.P. Bartel, P.L. Luisi, Synthesizing life, Nature 2001, 409, 387-390. 57741 [135] P.L. Luisi, T.P. de Souza, P. Stano, Vesicle Behavior: In Search of Explanations, Journal of 58742 Physical Chemistry B 2008, 112, 14655-14664. 59743 [136] J.E. Rothman, F.T. Wieland, Protein sorting by transport vesicles, Science 1996, 272, 227-234. 60 61 62 63 64 25

65 Directed Growth of Biomimetic 1 Microcompartments 2 3744 [137] F. Caschera, T. Sunami, T. Matsuura, H. Suzuki, M.M. Hanczyc, T. Yomo, Programmed 4745 Vesicle Fusion Triggers Gene Expression, Langmuir 2011, 27, 13082-13090. 5 6746 [138] K.P. Browne, D.A. Walker, K.J.M. Bishop, B.A. Grzybowski, Self-Division of Macroscopic 7747 Droplets: Partitioning of Nanosized Cargo into Nanoscale Micelles, Angewandte Chemie- 8748 International Edition 2010, 49, 6756-6759. 9749 [139] F. Caschera, S. Rasmussen, M.M. Hanczyc, An Oil Droplet Division-Fusion Cycle, 10 11750 Chempluschem 2013, 78, 52-54. 12751 13 14 15752 16 17753 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 61 62 63 64 26

65