Article Title: Engineering spatiotemporal organization and dynamics in synthetic cells

Article Type: Advanced Review

Authors:

First author

Alessandro Groaz, University of Michigan, [email protected]

Second author

Hossein Moghimianavval, University of Michigan, [email protected]

Third author

Franco Tavella, University of Michigan, [email protected]

Fourth author

Tobias W. Giessen, ORCID: 0000-0001-6328-2031, University of Michigan Medical School, [email protected]

Fifth author

Anthony G. Vecchiarelli, ORCID: 0000-0002-6198-3245, University of Michigan, [email protected]

Sixth author

Qiong Yang, ORCID: 0000-0002-2442-2094, University of Michigan, [email protected]

1

Seventh author

Allen P. Liu*, ORCID: 0000-0002-0309-7018, University of Michigan, [email protected]

Abstract

Constructing synthetic cells has recently become an appealing area of research. Decades of research in biochemistry and have amassed detailed part lists of components involved in various cellular processes. Nevertheless, recreating any cellular process in vitro in cell-sized compartments remains ambitious and challenging. Two broad features or principles are key to the development of synthetic cells - compartmentalization and self-organization/spatiotemporal dynamics. In this review article, we discuss the current state of the art and research trends in the engineering of synthetic cell membranes, development of internal compartmentalization, reconstitution of self-organizing dynamics, and integration of activities across scales of space and time. We also identify some research areas that could play a major role in advancing the impact and utility of engineered synthetic cells.

Graphical/Visual Abstract and Caption

Graphical Abstract. Self-organizing systems and dynamics in space and time in synthetic cells.

2

1. INTRODUCTION

As James Danielli first envisioned in 1972, biology, like all sciences, is going through an evolutionary process consisting of three phases, namely the observation phase, the analytical phase and the synthetic phase (Danielli 1972). We are currently witnessing the peak of the analytical phase of biology, as we continue to unravel cellular mechanisms that still have not been characterized. Nevertheless, like physics and chemistry before it, biology is starting to transition to its synthetic phase, as in the last decades we discovered and optimized methodologies for synthesizing DNA, RNA and proteins. In this regard, synthetic cells (also known as artificial cells) represent the natural evolution of biology as a science. Currently, synthetic cells are defined as encapsulated biochemical systems that can recapitulate one or a few of the functions that a natural cell can perform (Xu, Hu, and Chen 2016; Noireaux and Liu 2020) (Fig. 1). The engineering of synthetic cells can have a variety of purposes. Simple synthetic cells (often known as “protocells”), composed of chemically simple fatty acid membranes enclosing primitive biochemical processes, can be designed to probe the mechanisms of the origin of life on prebiotic Earth (Oberholzer et al. 1995; I. A. Chen, Salehi-Ashtiani, and Szostak 2005). Alternatively, synthetic cells can be designed to recreate specific cellular functions, thus enabling the study of discrete mechanisms or properties outside the complexity of a cell. Finally, although the technology is still in its infancy, synthetic cells hold great promise in biomedicine and biotechnology, they can be engineered to release molecules (e.g. insulin) in response to external stimuli (e.g. presence of glucose (Z. Chen et al. 2018)), or acting as programmable nano factories of bioactive drugs or as replacements of natural cells (Majumder and Liu 2017; T. Einfalt et al. 2018).

2. THE BOUNDARY OF SYNTHETIC CELLS

Simply put, a cell is not a cell without a membrane. A membrane boundary is necessary to 1) separate the cellular materials from the external environment and 2) permit selective communication of a cell with its surroundings via membrane-resident channels and transporters. In this section, we briefly discuss the importance of synthetic cell boundaries, the type of molecules used to build the boundary, and efforts in functionalizing this boundary.

2.1 The importance of a boundary in synthetic cells

3

Physically, the presence of a boundary of a synthetic cell permits the sensing of various mechanical stimuli, including tension, fluid shear or osmotic stresses. Biochemically, the cellular membrane acts as a semi-permeable barrier, permitting non-selective passive diffusion of some molecules, but also allows selective passive or active diffusion by using protein transporters and ion channels. In this way, biochemical reactions inside a cell are regulated by allowing the exchange of chemical information in and out of a cell. Nevertheless, the vast majority of molecules cannot cross the amphiphilic membrane. The diffusion across membranes has been exploited in artificial cell research to allow the bidirectional flow of chemicals. For instance, artificial cells consisting of an encapsulated cell-free expression reaction and genetic circuits have been engineered to uptake chemical messages that E. coli typically cannot sense and translate them into signals that are released via a synthesized pore protein to induce a response in E. coli (Lentini et al. 2014). This highlights the crucial role of a compartment to maintain concentrated transcription and translation machineries while actively exchanging messaging molecules with the external environment. However, relying on non-specific pore proteins or passive diffusion across the bilayer is often a limiting factor in synthetic cell communication applications. Ideally, only effector molecules should be allowed to diffuse out of the synthetic compartment, while other intermediates and substrates for transcription and translation should be retained in order to prevent premature cessation of protein synthesis.

Another crucial feature that a membrane can exhibit is the non-permeability to ions. The ability to segregate different ions on the inner and outer surfaces of a membrane creates an electrochemical gradient. The capability to generate, maintain and modify potential differences through the action of specialized membrane proteins is the core characteristic of excitable cells. Although to-date largely ignored, harnessing membrane potentials in synthetic cell research is a promising endeavor that could lead to a better understanding of electrical intra- and inter-cellular communications.

2.2 Building blocks for a synthetic cell boundary

A variety of self-assembling building blocks have been used as the chassis of synthetic cell boundaries (Fig. 2). Most notably, phospholipids, polypeptides, and block copolymers have been utilized to construct the membrane of synthetic cells. As the main building blocks of natural cell membranes, phospholipid bilayers maintain natural permeability as well as natural mechanical (e.g. elastic and flexible) and physical (e.g. fluid) properties. It is also the native substrate for hosting membrane proteins. Although microfluidic technologies have greatly facilitated the generation of phospholipid-

4

bound synthetic cells (Trantidou et al. 2018; Majumder, Wubshet, and Liu 2019), cell-sized liposomes remain mechanically fragile, especially when encapsulating cell-free expression systems.

Another type of self-assembling molecule that has gained interest recently is polypeptide copolymers. One example is elastin-like protein (ELP), which is an amphiphilic polypeptide that can self-assemble to compartmentalize biochemical reactions. Membranes of synthetic cells made of ELPs have tunable properties as these genetically encoded polypeptides have high chemical diversity (W. Kim and Chaikof 2010). Through facile cell-free synthesis of ELPs inside an ELP-based vesicle, a synthetic cell with the ability to grow in size was realized, demonstrating the possibility to use ELPs as a membrane constituent of synthetic cells (Vogele et al. 2018). In another study, vesicles of different sizes were constructed by diblocks of polypeptides that are responsive to external chemical stimuli (Bellomo et al. 2004). However, the polypeptide hydrophilic block from Bellomo et al. was composed of ethylene glycol-modified lysine residues. The use of modified amino acids makes it difficult to use a natural protein synthesis approach, and the diblock chain length flexibility was also quite limited. The resulting vesicles were responsive to stimuli only in high pH that makes them impractical in physiological conditions. Additionally, the reversible self-assembling property of ELPs as a function of their transition temperature provides the opportunity of constructing temperature-responsive synthetic cells (Martín et al. 2012). Nevertheless, ELPs are not the only peptides that can form enclosed structures that are able to host a biological system. For instance, surfactant proteins such as oleosins have been shown to self-assemble into micron-sized compartments whose physical characteristics can be tuned by engineering the sequence of the peptide (Vargo, Parthasarathy, and Hammer 2012). However, the structural tunability of oleosin polypeptides was limited to truncations from the N- and C- termini of proteins, and the effects of mutations in the sequence of the recombinant oleosin have not been extensively characterized.

Block copolymers represent the third group of molecules as building blocks of synthetic cell membranes. Block copolymers are made of a few domains of polymeric chains with variable lengths that show amphiphilic behavior. The advantage of block copolymer compartments, also known as polymersomes, over liposomes is their higher stability and stiffer structure. Depending on the structural design, polymersomes usually have thicker membranes that are less permeable compared to liposomes. As a step towards constructing multi-compartmentalized synthetic cells, micron-sized polymersomes encapsulating enzyme-filled nano-sized polymersomes can carry out an enzymatic cascade reaction (Peters et al. 2014). Even though the referenced study shows the possibility of mimicry of subcellular organization in a compartment, the practical application of this system as a platform for drug delivery in physiological conditions is limited. Illustration of enzymatic cascade reactions in compartments made of

5

biocompatible materials such as lipids or proteins can be envisioned as the next steps. Besides, the transport between synthetic organelles in this study relies solely on chemical diffusion that may not be optimal in more complex enzymatic reactions. Similarly, multi-compartmentalization of liposomes in polymersomes as a protected environment resembles natural organelles in living cells (Peyret et al. 2017). Combining the natural properties of phospholipids with the stability of block copolymers brings the possibility of hybrid membranes (Schulz and Binder 2015). Another example of a hybrid membrane employed organoclay/DNA that self-assembles into a giant compartment that can withstand the nucleation and growth of oxygen bubbles without breaking (Kumar, Patil, and Mann 2018).

The possibility to build hybrid membranes to compartmentalize enzymatic or cell-free expression reactions will allow the generation of synthetic cells with tunable biophysical and mechanical properties. The inclusion of compatible lipids or polymers in the hybrid membrane will likely allow the incorporation of membrane proteins in the envelope, vastly increasing the versatility of such synthetic cells. The consideration of components of membrane is critical given a major interest in synthetic cell engineering is to create a synthetic cell with the ability to grow and divide robustly in a controlled fashion.

2.3 Adding membrane proteins to synthetic cells

In natural cells, a plethora of membrane proteins carry out essential cellular functions ranging from transportation of nutrients to sensing of physical and chemical information. The majority of biological functionalities that have been recreated in synthetic cells have relied on different soluble enzymes or structural proteins that are encapsulated. The potential of synthetic cells for both basic research and translational applications could be greatly expanded by enriching the membrane with membrane proteins having specific functions. The incorporation of membrane proteins in synthetic cells is not trivial, but there have been a number of successes in recent years from either using purified or cell-free expressed proteins.

When it comes to communication across the membrane, the use of a non-specific pore alpha-hemolysin, either constitutively present or under genetic circuit control, has been a popular choice (Dwidar et al. 2019; Hilburger et al. 2019) since its pore size of 14 Angstrom allows for the retention of genetic material and most proteins yet allowing small molecules to pass through. Most importantly, it has a strong tendency to insert into lipid membranes upon reconstitution or by cell-free expression.

6

Among the most significant recent works based on reconstituting purified membrane proteins are a photosynthetic cell that couples light stimulation to drive ATP production for stimulating actin assembly (K. Y. Lee et al. 2018) or for driving protein synthesis (Berhanu, Ueda, and Kuruma 2019) and the reconstitution of integrin into vesicles for fibrinogen binding (Weiss et al. 2018). An emerging paradigm for membrane engineering relies on the synthesis and incorporation of membrane proteins in situ by the means of a cell-free expression system. Both prokaryotic and mammalian cell-free expression systems have been shown to successfully express structurally intact membrane proteins in mixed micelles or lipid bilayers (Shinoda et al. 2016; Majumder et al. 2019). Also, it has been demonstrated that mechanosensitive channel proteins can be successfully expressed and incorporated into synthetic cells (Majumder et al. 2017; Garamella et al. 2019; Hindley et al. 2019). Nevertheless, this approach is still in the early stages and only a few membrane proteins have been successfully expressed by prokaryotic and eukaryotic transcription-translation (TX-TL) systems and used in the membrane of synthetic cells to date.

Although lipid bilayers are the natural substrate for membrane proteins, synthetic cells with membranes of block copolymers or hybrid membrane hosting functional membrane proteins or pores have been constructed (H. J. Choi and Montemagno 2007; Tomaž Einfalt et al. 2015; Jacobs, Boyd, and Kamat 2019). However, the effect of polymer properties including the length of each block on the membrane protein functionality or its folding characteristics has not been thoroughly investigated. It is worth noting that the translocation of membrane proteins into the polymer or hybrid membranes, in general, may not necessarily be better than phospholipid bilayer membranes.(Göpfrich, Platzman, and Spatz 2018; Schwille et al. 2018; Wang et al. 2020) In another work, the bottom-up reconstitution of an artificial mitochondrion was demonstrated in both polymersomes and hybrid membrane vesicles by reconstituting an ATP synthase and a terminal oxidase, furthering the possibility of using hybrid membranes in synthetic cells (Otrin et al. 2017).

2.4 Perspectives on the engineering of synthetic cell membranes

Synthetic cell research could be greatly advanced by engineering new functionalities at the synthetic cell boundary. For instance, generalized methods for the incorporation of in situ cell-free expressed adhesion proteins, ion channels, transporters and signaling receptors would allow sensing of different stimuli, e.g. light, ligands, or voltage. Combined with biochemical or genetic circuits, we envision the reconstitution of signaling pathways in synthetic cells. Synthetic cells with an ‘active’ membrane could enable the study of mutant membrane proteins outside of the complexity of a cell. If successfully

7

developed, a synthetic cell platform for membrane protein reconstitution would be much simpler compared to conventional membrane protein purification and would be amenable for drug screening studies as well in building smart drug delivery vehicles.

3. COMPARTMENTALIZATION OF SPACES

Cellular compartmentalization and spatial control are defining features of life (Diekmann and Pereira- Leal 2013). Organelles exist to compartmentalize functions in a cell. In this section, we discuss emerging concepts of cellular compartmentalization including several protein-based organelles found in bacteria and biomolecular condensates. We will not discuss lipid- and vesicle-based strategies explicitly in this section as these approaches are more established and already covered in a number of excellent recent review articles (Göpfrich, Platzman, and Spatz 2018; Schwille et al. 2018; Wang et al. 2020).

3.1 Self-assembling protein organelles

Similar to eukaryotes, many prokaryotes compartmentalize their cytosol to carry out specialized metabolic reactions, prevent toxicity and store nutrients (Cornejo, Abreu, and Komeili 2014; Martin 2010). To achieve this goal, prokaryotes rely on protein-based instead of membrane-based strategies (Figure 3A) (Giessen and Silver 2016a; Kerfeld et al. 2018). Protein organelles are functional analogues of eukaryotic membrane organelles. A semipermeable protein shell creates a sequestered reaction space separated from the bulk cytosol (Kerfeld and Erbilgin 2015). This compartmentalization strategy can increase the local concentrations of enzymes and metabolites (Iancu et al. 2010; Ting et al. 2015), prevent the leakage of toxic or volatile intermediates (Bobik, Lehman, and Yeates 2015) and create unique microenvironments with regard to pH, redox state and cofactor pools (Figure 3B and 3C) (Kerfeld and Melnicki 2016). Protein organelles enable specialized biochemistry that would not be possible without compartmentalization and are so far known to be involved in carbon fixation (Kerfeld and Melnicki 2016), carbon source utilization (Axen, Erbilgin, and Kerfeld 2014), iron metabolism (Bobik, Lehman, and Yeates 2015) and stress resistance (Nichols et al. 2017). The two main classes of microbial protein organelles are encapsulin nanocompartments and bacterial microcompartments (BMCs) (Bobik, Lehman, and Yeates 2015; Nichols et al. 2017; Kerfeld et al. 2018). In both cases, specialized enzymatic machinery is selectively encapsulated within a self-assembling protein shell. Genes encoding shell proteins and the enzymatic core cluster together and are organized in co-regulated operons (Axen, Erbilgin, and Kerfeld 2014; Bobik, Lehman, and Yeates 2015; Giessen 2016). The fact that protein

8

organelle structure and function are genetically encoded, which is not the case for lipid-based organelles, makes them highly engineerable and promising starting points to build designed compartmentalization systems in a synthetic cell context (Giessen and Silver 2016a).

3.2 Bacterial microcompartments (BMCs)

BMC operons are widely distributed in bacteria and can be found in the majority of bacterial phyla (Axen, Erbilgin, and Kerfeld 2014). BMCs assemble into large cytosolic protein compartments with diameters between 40 and 500 nm composed of a selectively permeable protein shell and an enzymatic core (Tanaka et al. 2008; Cléement Aussignargues et al. 2015; Sutter et al. 2017). BMC shells generally consist of 3 types of small proteins that form hexameric, pseudo-hexameric and pentameric complexes representing individual facets which self-assemble to form the overall BMC shell. BMCs specifically sequester specialized enzymes and are involved in the anabolic fixation of carbon (carboxysomes) and catabolic processes like carbon and nitrogen source utilization (e.g. 1,2-propanediol (PDU), ethanolamine (EUT), choline, l-fucose) (Petit et al. 2013; Turmo, Gonzalez-Esquer, and Kerfeld 2017; Kerfeld 2017; Herring et al. 2018). All known catabolic BMCs, also known as metabolosomes, encapsulate enzymatic pathways that proceed through toxic aldehyde intermediates and it has been proposed that the compartment shell protects the rest of the cell from these toxic intermediates (Chowdhury et al. 2014). Metabolosomes have been suggested to provide bacteria with competitive advantages in complex environments and have been shown to be important for the virulence of some widespread pathogens like Salmonella typhimurium (Klumpp and Fuchs 2007; Harvey et al. 2011; Thiennimitr et al. 2011) and Clostridium difficile (Pitts et al. 2012; Bhardwaj and Somvanshi 2017). Recent bioinformatic searches have revealed dozens of uncharacterized BMC systems with unknown functions encoded in human-associated bacteria (Axen, Erbilgin, and Kerfeld 2014).

BMCs have recently attracted the attention of protein engineers and synthetic biologists as modular engineerable systems with which to control protein and even pathway compartmentalization with the ultimate goal of building designer nano-bioreactors (Plegaria and Kerfeld 2018; Planamente and Frank 2019; M. J. Lee, Palmer, and Warren 2019). So far, most studies aimed at engineering BMCs have been carried out in vivo. One important prerequisite for BMCs to be useful tools in cell-free and synthetic cell settings is the ability to decouple the formation of separated spaces, i.e. shell formation, from the natively associated enzymatic cargo. This has been achieved through the co-expression of shell components for a number of systems including PDU, EUT and b-carboxysome BMCs resulting in the in vivo assembly of empty BMC-like shells (Parsons et al. 2010; Choudhary et al. 2012; Lassila et al. 2014;

9

Mayer et al. 2016; Cai et al. 2016). By combining shell components from different BMC systems, it is also possible to form chimeric BMC shells (Cai et al. 2015; Fang et al. 2018).

The second prerequisite needed to make BMCs useful tools for synthetic biology is a modular targeting mechanism for selectively localizing arbitrary protein components inside a BMC shell (E. Y. Kim and Tullman-Ercek 2014). Non-native enzymes can be directed to the BMC interior by genetic fusion of so- called encapsulation peptides (EPs) found either at the N- or C-termini or internal regions of an enzymatic core component (Jakobson et al. 2015). EPs are generally amphipathic helices and EPs from different BMC systems can often be used interchangeably due to a conserved interaction mechanism. The main drawbacks of using native EPs to target heterologous cargo are their low targeting efficiency and lack of control over stoichiometry (E. Y. Kim and Tullman-Ercek 2014). This has led to the development of alternative targeting strategies, including rational design and library-based screening of de novo EPs, orthogonal covalent systems like the SpyTag/SpyCatcher targeting system based on split bacterial adhesin domains, and split inteins (Zakeri et al. 2012; Y. Li 2015; Hagen, Sutter, et al. 2018).

A third prerequisite that needs to be considered is gaining control over BMC shell permeability which controls matter and information exchange between the BMC interior and the bulk cytosol. All BMCs possess distinct pores at the center of their (pseudo)-hexameric and pentameric facets. Two approaches have been used to modulate the permeability of BMCs. First, the residues lining the pores can be altered to change the size, charge or hydrophobicity of a given pore (Slininger Lee, Jakobson, and Tullman-Ercek 2017). Second, chimeric facets can be created by combining shell protein monomers from different BMC systems (Cai et al. 2015). Additionally, de novo designed iron-sulfur cluster binding sites have been engineered in heterologous BMC pores which may allow future control over electron transfer into and out of BMCs (Clément Aussignargues et al. 2016). Applying the principles discussed above, BMCs have been repurposed for a number of novel functions including the encapsulation of a non-native two- enzyme metabolic pathway for increased ethanol production in a heterologous host (Lawrence et al. 2014). Engineered BMCs have further been employed for polyphosphate accumulation and enhanced safe expression of toxic proteins (Liang et al. 2017; Yung et al. 2017).

While BMC engineering has been primarily carried out in vivo, some advances have been made on in vitro assembly. A major issue for in vitro reconstitution of BMC shells is the tendency of all individual shell components to self-associate, preventing their easy purification. A number of studies were able to counteract self-association via fusion tags (His6 and SUMO) which allowed purification of individual shell proteins (Lassila et al. 2014; Uddin et al. 2018; Hagen, Plegaria, et al. 2018). Upon mixing them in

10

defined stoichiometries followed by tag removal, BMC-like closed shells as well as protein nanotubes could be formed. Notably, in vitro assembled BMC shells are generally smaller than native BMCs, indicating that cargo proteins are important for determining shell dimensions. The importance of relative protein ratios and stoichiometry is further highlighted by the fact that many higher-order and extended architectures can be observed when heterologously expressing BMCs in vivo or assembling them in vitro. These include flat sheets, tubes, filaments and swiss-roll like structures (Dryden et al. 2009; Heldt et al. 2009; Choudhary et al. 2012).

Currently, a number of major challenges exist which preclude the easy use of BMCs in cell-free contexts. BMCs are many megadalton in size and their shells are often made up of between 3 and sometimes more than 10 components. Formation of closed BMC-like shells relies on the precise control over relative protein ratios, otherwise aberrant and non-closed structures will be formed. This can be addressed by developing cell-free systems where the relative expression levels of many components (shell and cargo) can be precisely controlled. Another useful future strategy will be the removal of redundant shell components to build more streamlined minimal shell systems. This will allow the controlled assembly of very large shells acting as genetically encoded sub-compartments in cell-free systems and synthetic cells.

3.3 Encapsulin nanocompartments

Recent genome-mining studies have shown that encapsulin operons are widely distributed in bacterial and archaeal genomes (Giessen and Silver 2017a; Tracey et al. 2019). Encapsulins self-assemble into icosahedral protein compartments between 20 and 50 nm in diameter and are able to selectively encapsulate cargo enzymes involved in specialized metabolic functions (Akita et al. 2007; Sutter et al. 2008; McHugh et al. 2014). Encapsulin shells consist of a single type of shell protomer and cargo encapsulation is mediated by short C-terminal sequences found in all cargo proteins called targeting peptides (TPs) (Tamura et al. 2015; Cassidy-Amstutz et al. 2016). Encapsulin capsid proteins possess the HK97 phage-like fold and most likely originate from defective prophages whose capsid components have been co-opted by host metabolism (Giessen and Silver 2017a; Nichols et al. 2017). This molecular domestication of a viral capsid-forming shell protein by microbes represents an intriguing example of the evolutionary entanglement of viral and cellular protein folds. Encapsulin systems have been implicated in oxidative and nitrosative stress resistance, iron and redox homeostasis and the biosynthesis or degradation of so far uncharacterized metabolites (Rahmanpour and Bugg 2013; Contreras et al. 2014; Giessen and Silver 2017a; Nichols et al. 2017).

11

So far, no study producing encapsulins in a cell-free system has been published with most of the encapsulin engineering work having been done in vivo (Giessen 2016; Giessen and Silver 2017b). Similar to BMCs, encapsulins represent promising systems for controlled compartmentalization of proteins and enzymatic pathways as well as engineered spatial control at the molecular level. Targeting of non-native cargo proteins to the interior of encapsulins has been reported multiple times by genetic fusion of TPs to proteins of interest (Tamura et al. 2015; Cassidy-Amstutz et al. 2016; Lau et al. 2018; B. Choi et al. 2018; Künzle et al. 2018). Various non-native protein fusions to either the N- or C-terminus of the encapsulin capsid protein have been reported resulting in the encapsulation or external display of a fused protein of interest. These engineered encapsulin systems have been used in a number of biotechnological and biomedical applications including targeted drug delivery and synthesis (Moon et al. 2014; B. Choi et al. 2016; Sonotaki et al. 2017; Bae et al. 2018; T. H. Lee et al. 2020) biomaterials and nanoparticle synthesis (Giessen and Silver 2016b; Künzle et al. 2018), vaccine design (Lagoutte et al. 2018), nano-bioreactors for improved biocatalysis (Putri et al. 2017; Lau et al. 2018) and subcellular labelling reagents (Sigmund et al. 2018; 2019). Further, encapsulin pores, found at the 5- and 3-fold symmetry axes of the icosahedral shell, have been engineered to modulate shell permeability (Williams et al. 2018).

The fact that encapsulin shells are able to assemble into closed structures from a single type of shell protein makes it more straightforward to use them in in vitro settings compared to BMCs. Reversible in vitro assembly and disassembly of encapsulins has been achieved based on pH shifts (Cassidy-Amstutz et al. 2016). This is encouraging for potential future applications of encapsulins as dynamic stimulus- responsive compartmentalization systems in cell-free and synthetic cell systems. Even though encapsulin self-assembly in cell-free systems has not yet been reported, other protein compartments of similar size and complexity have been successfully expressed and assembled using bacterial cell-free systems. This includes the cell-free production of T7, ΦX174 and MS2 bacteriophage capsids which share many features with encapsulin shells (Shin, Jardine, and Noireaux 2012; Garamella et al. 2016).

Overall, encapsulins face fewer hurdles than BMCs with regard to their application in synthetic cells primarily due to their smaller size and more straightforward cargo loading and assembly pathways. Ideally, both encapsulins and BMCs will be used in parallel in future synthetic cell systems to provide genetically encoded and dynamic compartmentalization strategies that span several orders of magnitude of volume (20 to 500 nm in diameter).

3.4 Liquid-liquid phase separation and membraneless protein organelles

12

In recent years, there is mounting evidence that liquid-liquid phase separation (LLPS) represents a key cellular mechanism for the formation of membraneless protein compartments (Hyman, Weber, and Jülicher 2014). The liquid nature of protein- and RNA-rich compartments was first established in P granules found during germ cell specification in C. elegans (Brangwynne et al. 2009). The drivers of LLPS in many cases involve intrinsically disordered proteins (IDPs) or regions (Elbaum-Garfinkle et al. 2015). For instance, both LAF-1 and PGL-3, which are P granule proteins, contain arginine/glycine-rich (RGG) domains which are necessary and sufficient for LLPS (Elbaum-Garfinkle et al. 2015; Saha et al. 2016). LLPS observed in membraneless compartments has been recapitulated biochemically using purified proteins in vitro (P. Li et al. 2012; Molliex et al. 2015; Jain and Vale 2017; Schuster et al. 2018; Du and Chen 2018), making it another promising experimental platform to explore approaches for organizing intracellular activity in synthetic cells. Encapsulation of a two-phase system in a synthetic cell context was first demonstrated many years ago using dextran and PEG solutions where their mixing/demixing is controlled by concentration or temperature (Long et al. 2005). With increasing appreciation of LLPS in biological organization, recent studies have shown that formation of protein-rich condensates can support bacterial cytoskeleton dynamics or enzymatic polymerization of nucleic acids (Monterroso et al. 2019; Deshpande et al. 2019). Coacervate droplets have been repurposed as artificial organelles for spatial organization of bio-reactions and incorporating LLPS as artificial organelles in synthetic cells is a frontier of current synthetic cell research (Deng and Huck 2017; Crowe and Keating 2018). Recently, protease-triggered assembly and disassembly of RGG-based protein droplets has been demonstrated as well (Schuster et al. 2018). Other recent examples of dynamically controlling compartmentalization in cell-free liposome-based systems rely on the pH-responsiveness of polylysine/adenosine triphosphate or spermine/RNA to dynamically form single coacervate droplets within lipid vesicles (Last, Deshpande, and Dekker 2020). In one study, the enzyme formate dehydrogenase was shown to partition and concentrate within these liquid-liquid phase separated droplets which led to a concentration-based activation of enzyme activity (Love et al. 2020).

By integrating protein shell- and liquid condensate-based approaches for the spatial and temporal control of biochemical reaction networks in synthetic cells (Figure 3D), it will be possible in the future to create a general platform technology adaptable to investigate and engineer many aspects of synthetic cell metabolism. This will not only allow novel insights into how compartmentalized spaces at the nanoscale facilitate the proper functioning of natural cells, but also the creation of programmable synthetic cells for applications as nanoreactors and production platforms, biosensors, delivery devices and building blocks for biomimetic materials.

13

4. SELF-ORGANIZATION, POSITIONING, AND DYNAMICS

A fascinating feature of living cells is their ability to self-organize and position molecules in space and time. In the following section, we describe several reconstitution systems related to bacterial chromosome organization, segregation, and that are particularly attractive in the context of synthetic cell research.

4.1 Chromosomal organization

Bacterial chromosomes are typically orders of magnitude longer than their cellular confines. Physical and biochemical processes organize chromosomes into a compact, but accessible nucleoid (Jun 2015). Polymer dynamics of the chromosome provides ~ 100-fold compaction by molding it into a globule (Surovtsev and Jacobs-Wagner 2018). Cellular confinement, crowding, and excluded volume effects further aid in compaction (Murphy and Zimmerman 1997; Odijk 1998; Cunha, Woldringh, and Odijk 2001). DNA supercoiling by topoisomerases is also essential in sustaining a compact yet accessible genome. Nucleoid-associated proteins (NAPs) are key to compacting and organizing bacterial chromosomes. NAPs are functionally analogous to eukaryotic histones, compacting the chromosome by kinking and bridging the DNA (Thompson and Landy 1988; Amit, Oppenheim, and Stavans 2003; Dame, Noom, and Wuite 2006). Structural maintenance of chromosome (SMC) complexes are also of critical importance, and likely function by co-entrapment of DNA loops within the circumference of the SMC ring (Gligoris et al. 2014). In many bacteria, a single SMC complex is sufficient for normal growth of cells (Hutchison et al. 2016).

The molecular mechanism underlying SMC-based chromosome compaction has remained elusive. Recently, the Dekker group provided direct visualization of an SMC complex forming and processively extruding DNA loops in a cell-free setup using purified components (Ganji et al. 2018; E. Kim et al. 2020). These studies are significant steps forward in the bottom-up reconstitution of chromosome compaction. An attractive next step would be to combine circular DNA molecules with SMC complexes and topoisomerases within liposomes. Such experiments would determine if physical processes (polymer dynamics, confinement, crowding) combined with supercoiling and SMC-based looping are necessary and sufficient in the formation of a minimal nucleoid.

4.2 DNA segregation

14

Prior to cell division, replicated DNA must be segregated and positioned to opposite sides of the cell to ensure faithful genetic inheritance. Many bacterial chromosomes and plasmids employ partition (Par) systems. Par systems are minimal and self-organizing, encoding a cis-acting DNA-binding site on the chromosome (or plasmid) and two trans-acting proteins (Baxter and Funnell 2014). The DNA-binding site is bound by one protein, while the other protein is an NTPase that uses ATP or GTP hydrolysis to drive the segregation reaction. Par systems are grouped according to the NTPase type - whether it contains a Walker ATP-binding motif (ParA), or is structurally similar to eukaryotic actin (such as ParM) or tubulin (such as TubZ) (Gerdes, Howard, and Szardenings 2010). All three types have been reconstituted and imaged in cell-free setups using purified and fluorescently labeled components.

The ParM system and its polymer-based mechanism for segregating DNA is well understood because of cell-free reconstitution data that correlates strongly with in vivo function and dynamics. Through a mechanism of insertional polymerization, ParM filaments push plasmids to opposite cell poles in vivo (Møller-Jensen et al. 2003). In a cell-free setup, ParM undergoes ATP-dependent polymerization; pushing apart beads coated with the partner protein (ParR) bound to its cognate DNA-binding site (parC) (Garner, Campbell, and Mullins 2004; Garner et al. 2007). TubZ polymers have also been reconstituted (Fink and Löwe 2015) and show dynamic instability (Erb et al. 2014) and treadmilling (Larsen et al. 2007) in the presence of GTP.

ParA-based systems are encoded by 70% of all sequenced bacterial chromosomes and also found on most low copy plasmids (Baxter and Funnell 2014). Studies on plasmid partitioning have been particularly useful in elucidating the ParA-based mechanism. In vivo, the ParA ATPase coats the nucleoid, while its partner protein, ParB, binds and spreads from its DNA-binding site (parS) on the plasmid. The ParB-parS complex stimulates the release of ParA proteins from the nucleoid in the vicinity of the plasmid, resulting in a concentration gradient of ParA. Following replication, plasmid copies segregate as they chase the ParA gradient in opposite directions. This gradient-based mechanism ensures that each daughter cell inherits the plasmid after division. Gradient-based mechanisms for ParA-based chromosome segregation have also been proposed (Lim et al. 2014; Le Gall et al. 2016; Walter et al. 2017).

The Mizuuchi group reconstituted ParA-based DNA segregation, movement, and positioning using purified and fluorescently labeled components imaged on a DNA-carpeted flow cell, which served as a biomimetic of the nucleoid (Hwang et al. 2013; Vecchiarelli, Hwang, and Mizuuchi 2013; Vecchiarelli, Neuman, and Mizuuchi 2014). Purified and fluorescently labeled variants of ParA and ParB were mixed

15

with either plasmid or DNA-coated beads containing the parS binding site. Similar to in vivo dynamics, ParA coated the DNA carpet and its concentration was depleted in the vicinity of ParB-bound plasmid or bead substrates. These “DNA-cargoes” utilized the ParA gradient on the DNA carpet for directed movements.

4.3 Positioning cell division

Cell division machinery in bacteria can be positioned by several positive and negative regulators (Monahan et al. 2014). But to date, only the Min/FtsZ system has been reconstituted. The Min system acts on a tubulin-like GTPase called FtsZ; a highly conserved protein among bacteria and also found in chloroplasts and mitochondria (Leger et al. 2015; C. Chen et al. 2018). FtsZ is the most extensively studied protein involved in the division of a bacterial cell (den Blaauwen, Hamoen, and Levin 2017). FtsZ polymerizes into a structure called the Z ring. The Z ring (i) acts as a scaffold for the recruitment and assembly of several other division proteins, (ii) contributes to the invagination force, and (iii) organizes cell wall remodeling during septation (den Blaauwen and Luirink 2019). FtsZ does not directly bind to the membrane but depends on adaptor proteins such as FtsA and ZipA. In the presence of GTP, treadmilling FtsZ polymers have been reconstituted on supported lipid bilayers (SLBs) using its native membrane anchors FtsA and ZipA (Pazos, Natale, and Vicente 2013; Rico, Krupka, and Vicente 2013). FtsZ and FtsA have also been shown to form helical co-polymers that mechanically constrict liposomes and generate narrow necks, but complete division did not occur (Szwedziak et al. 2014). Very recently, FtsA-FtsZ ring- like structures were reconstituted by way of a cell-free gene expression system both on planar membranes and inside liposome compartments (Godino et al. 2019, Godino et al. 2020). The ‘cytoskeletal structures’ were found to constrict the membrane and generate a few budding vesicles. How to faithfully generate robust division events via an FtsZ-based mechanism in a liposome remains unclear.

The Min system consists of three proteins that self-organize into a cell pole-to-cell pole oscillator on the inner membrane, which aligns Z-ring assembly at mid-cell by preventing FtsZ polymerization at the cell poles (Szwedziak and Ghosal 2017). MinD, when bound to ATP, associates with the inner membrane (Hu and Lutkenhaus 2003; Zhou and Lutkenhaus 2003). MinE also has membrane-binding activity and can interact with membrane-bound MinD (Hu, Gogol, and Lutkenhaus 2002; Vecchiarelli et al. 2016). MinE interactions with MinD have been shown to recruit, stabilize, release, and inhibit MinD interactions in the membrane (Mizuuchi and Vecchiarelli 2018). This dynamic interplay between MinD and MinE results

16

in a pole-to-pole oscillation on the membrane (Raskin and De Boer 1999). The third protein, MinC, is not required for oscillation, but is the inhibitor of Z-ring assembly (Lutkenhaus 2007). MinC acts as a passenger protein, associating with MinD on the membrane and inhibiting FtsZ polymerization. The pole-to-pole oscillation of MinD therefore recruits the inhibitory activity of MinC at the poles, which promotes Z-ring formation and symmetric cell division at mid-cell.

The Min system has been shown to form a wide-variety of patterns when unleashed from cellular confines (Fig. 4A) and reconstituted in a variety of cell-free formats (Loose et al. 2008; 2011; Schweizer et al. 2012; Petrasek, and Schwille 2013; Zieske and Schwille 2014; Vecchiarelli et al. 2014; Vecchiarelli et al. 2016; Caspi and Dekker 2016; Kretschmer, Zieske, and Schwille 2017; Ramm, Glock, and Schwille 2018; Martos, Kohyama et al. 2019): (i) open-well and flow-channel microfluidics, (ii) E.coli extracts and synthetic lipid mixtures of defined composition, (iii) SLBs and liposomes, and (iv) various 2D and 3D confinement geometries. In these setups, the biochemistry driving Min patterning can be visualized. MinD and MinE mutants, changes in protein concentration and stoichiometry, and changes in buffer conditions (i.e. salt concentration, molecular crowding) have allowed for the development of comprehensive molecular mechanisms.

The Schwille group has reconstituted the entire Min/FtsZ system in rod-shaped, lipid-coated troughs that mimic the geometry of an E. coli cell (Zieske and Schwille 2014). In this cell-free setup, a fluorescent version of FtsZ was fused to a membrane targeting sequence (FtsZ-YFP-mts), which allowed FtsZ to directly bind membrane, bypassing the need for membrane recruitment by FtsA and ZipA. Strikingly, the pole-to-pole oscillation of the MinCDE system restricted Z-ring assembly to the mid-point of the rod- shaped trough (Fig. 4B). The reconstitution of the Min oscillator inside free-standing lipid vesicles was recently achieved (Litschel et al. 2018; Godino et al. 2019). Litschel et al. encapsulated purified MinE and fluorescent MinD into vesicles 5 to 100 microns in diameter, which allowed for precise control of the relative and absolute concentrations of Min components. Godino et al., on the other hand, synthesized MinD and MinE de novo in vesicles 1 to 20 microns in diameter using the PUREfrex2.0 TX-TL system. A purified fluorescent fusion of the MinC protein was also added at a fixed concentration. As a result, the absolute and relative concentrations of all three proteins were highly variable for experiment-to- experiment. But in both cases, the Min system was observed to form patterns that are similar to those formed on SLBs: pulsing vesicles (global binding and release from the membrane), pole-to-pole oscillation, circling waves, and trigger waves. Intriguingly, vesicle shape oscillations were also observed, and in some cases, these deformations led to periodic vesicle splitting. This finding suggests that it may be possible to divide a liposome-based synthetic cell via “budding” induced by Min oscillations, and

17

without FtsZ. Godino and colleagues, in a separate publication, were also recently successful in cell-free gene expression of FtsZ-FtsA ring-like structures that could constrict liposomes; generating elongated membrane necks and, on occasion, budding vesicles (Godino et al. 2020). To date, however, these stepwise additions of divisome components have yet to show consistent abscission of liposomes. The next important step is to combine the Min/FtsZ system (MinCDE with FtsZ and one or more of its membrane anchors such as FtsA) into a liposome. Additional downstream division proteins are likely required for reconstituting an FtsZ-based division machine from scratch.

5. SCALES IN TIME AND SPACE

Cells operate over an astounding range of scales in both time and space. From nanoscale enzymes that catalyze reactions in microseconds to micron-sized cells that divide in tens of minutes, cellular life is as versatile as it is complex. Therefore, designing synthetic cells poses the unique challenge of reproducing processes at this breadth of spatio-temporal scales. In what follows, we will review the recent achievements in the reconstitution of cellular processes and point out how they contribute to achieving a multiscale functioning synthetic cell. Then, we will discuss the integration of these processes and the crucial role of modeling in the endeavor of engineering synthetic cells.

5.1 Time

Cells are intrinsically dynamic across a large range of temporal scales (Fig. 5). On the fast end (fractions of a second to a few minutes), macromolecular interactions and post-translational modifications serve as a platform for cellular signaling. On the slow end (hours to days), biomolecular networks, including gene regulation networks and metabolic networks, coordinate to fulfill essential functions like cell cycles, metabolism, and multicellular aggregation. A key challenge in designing a comprehensive synthetic cell is to construct self-organizing processes that couple both slow and fast timescales to provide versatile while accurate performance. Over the last decade, scientists have begun to achieve the reproduction of certain biological processes spanning a considerable range of cellular dynamics (Fig. 5). Below, we will review recent synthetic cell research focusing on cell cycle and division, metabolism, multicellularity, and scalability.

The cell cycle is a fundamental self-organizing process that governs when a cell is to divide. The temporal length between two successive divisions can vary from tens of minutes in bacteria and some eukaryotic cells including yeast and early embryonic cells, to about a day in somatic cells. Unlike binary 18

fission of how most prokaryotes divide, eukaryotic cell cycle is regulated by a more sophisticated regulatory network that may involve hundreds of genes, proteins, and signaling molecules (e.g. KEGG human cell cycle pathway: hsa04110), making it difficult to reproduce the process from scratch. Different strategies have been applied to mimic both prokaryotic and eukaryotic cell cycle oscillations and divisions in synthetic cells. As aforementioned in section 4.3, using a cell-free TX-TL system known as PURE (Shimizu et al. 2001), a Min oscillatory system has been integrated with the cytoskeletal proteins to dynamically regulate FtsZ patterns, as a way to regulate the prokaryotic cell division in vitro (Godino et al. 2019). The oscillation, at a period of 25 seconds, was observed for up to 6 hours. Utilizing Xenopus laevis egg with recombinant proteins, mRNAs, and DNA, a self-sustained eukaryotic cell cycle has also been reconstituted in cell-sized water-in-oil microemulsion droplets in both cytosolic- only and nuclei-assembled forms (Guan, Wang, et al. 2018; Guan, Li, et al. 2018; Sun et al. 2019). The cell cycles in microdroplets featured a period tunable from tens to hundreds of minutes and lasted for several days. Additionally, in a 2D chamber environment, the cycling egg cytoplasm supplied with sperm DNA self-organizes into cell-like compartments that display mitotic divisions at each cell cycle (Cheng and Ferrell 2019). These cell-like in vitro microenvironments, comparing to biological cells, have a great flexibility in circuit dissection, speed tuning, and manipulation of concentration, stoichiometry, size, nucleus-to-cytoplasm ratio, energy, etc., providing a powerful platform for quantitative characterization of cell-cycle machinery. Despite these advantages, most cell-cycle reconstructions so far relied on cell- extracted materials (e.g. cytosol, sperm chromatin, etc.) that are still too complex to gain insight into a minimal design for the self-organizing process. An interesting next step in bottom-up synthetic biology is to reconstruct a cell cycle completely from scratch. Computational studies have shown that even with complicated biological clock networks, a minimal core architecture is sufficient for self-sustained oscillations, while those evolutionary-conserved peripheral network structures may provide additional modifications to promote functions such as robustness (Z. Li, Liu, and Yang 2017) and tunability (Tsai et al. 2008). Indeed, a landmark study has successfully reconstituted a temperature-compensated bacterial circadian clock in a test tube with components as simple as three Kai proteins and ATP (Nakajima et al. 2005), exemplifying a minimal design of a circadian clock. Constructing a minimal cell-cycle circuit from the bottom-up, for instance, starting from a set of hypothetical central mitotic regulators including cyclin B1, cyclin dependent kinase 1 (Cdk1), and anaphase promoting complex/cyclosome (APC/C), can not only test our existing understanding of the core architecture but also serve as a foundation with auxiliary structures added one at a time, a step forward towards a synthetic cell with well-predicted cell- cycle functions. A foreseeable challenge of building a bottom-up out-of-equilibrium system like cell cycles is to self-sustain energy. Incorporation of metabolic cycles with cell cycles can be an additional

19

interesting goal in this direction. In the paragraph below, we will review the recent effort in building machinery to replenish energy as well as other consumable components.

Metabolism is another cornerstone process in living systems. The central challenge is to achieve bottom- up integration of sustainable metabolic functions in synthetic cells. Two important studies achieved this result: one using E. coli inverted membrane vesicles to create a minimal metabolism containing an NAD reaction and an NAD-regeneration module (Beneyton et al. 2018), and the other creating light-driven artificial organelles by proteoliposomes formed of bacteriorhodopsin and ATP synthase to allow photo- synthesizing ATP inside a giant unilamellar vesicle (GUV) (Berhanu, Ueda, and Kuruma 2019). Both systems were able to function autonomously for hours. The presence of metabolic activity is considered a key factor in discerning between living and non-living matter. Refilling energy and other cellular consumables (e.g. lipid membranes, see section 5.2) is important for an autonomously functioning and growing synthetic cell. Studies that have thus far implemented a few different sustaining machineries like NAD and photosynthetic pathways are the beginning steps by synthetic biologists into this exciting field. Moreover, the design of repair mechanisms as well as routes for the elimination of byproducts, appear to be future challenges in this area of research. Finally, greater integration between these cyclic biochemical processes and other physical and mechanical activities can serve as a way to further empowering a synthetic cell for controlling its own life cycle.

Besides reproducing intracellular processes, it is also important to emulate the higher-order intercellular interactions observed among biological cells. Applying engineered regulatory networks and synthetic signaling pathways, studies have achieved multicellularity or cell-cell communications through various strategies. Examples include programming synNotch signaling for cell adhesion and multicellular structures (Toda et al. 2018), engineering genetic circuit-containing synthetic minimal cells (synells) to enable liposome interaction and fusion (Adamala et al. 2017), creating quorum sensing machinery in phospholipid vesicle-based artificial cells (Lentini et al. 2017), and communicating non-lipid microcapsules through DNA strand displacement (Joesaar et al. 2019). However, all these processes need to be responsive to an ever-changing environment. Typically, cells sense inputs that are transduced by signaling pathways which ultimately change the expression level of a group of genes. Therefore, it is important to comprehensively study the connection between long-timescale responses and short-timescale input sensing. Using motifs called phospho-regulons to create a dual-timescale regulatory network, cell responses can be controlled and tuned over a wide dynamic range (Gordley et al. 2016). Despite efforts of the synthetic biology community to standardize biological parts, rational design of transcription factors remains challenging. To overcome this issue and to promote scalability

20

and modularity, a workflow for an exhaustive characterization of transcription factors was recently presented (Swank, Laohakunakorn, and Maerkl 2019) and an integrated regulatory network with multiple functional modules was created (Schaffter and Schulman 2019).

Different processes have been reproduced in vitro across the relevant timescales of cellular life (Fig. 5). However, in a biological cell all these processes need to run in parallel cooperatively. Therefore, moving forward, two avenues of research will probably thrive in the field of synthetic cells. On the one hand, new processes and modules will be built to enrich the existing arsenal of building blocks. On the other hand, the readily available building blocks will be integrated to achieve increasingly complex cellular behaviors. Importantly, the accomplishments in the time domain of synthetic cells have to be mirrored by similar developments in the spatial domain.

5.2 Space

From molecules (~nm) to developing eggs (~µm to mm), cellular life is a remarkable multi-scale system. Like temporal scales, spatial scales are crucial to cell functions. Dynamics in bulk, usually characteristic of ensemble averages, cannot represent behaviors in single cells. Moreover, certain processes including cell cycles (Jones et al. 2017; Guan, Li, et al. 2018) can progress in a size-dependent manner. Therefore, constructing a synthetic cell requires creating micrometric cellular and subcellular compartments filled with components compatible with cellular life (Fig. 6). One of the most prominent experimental techniques for achieving these scales is the encapsulation of cellular processes in water-in-oil emulsions and GUVs via microfluidics. Such a platform permits precision, reproducible and high-throughput generation of encapsulated systems at a wide range of cell-sized and subcellular scales.

The versatility of microfluidics allows the creation of micrometric chambers (Tayar et al. 2017; Swank, Laohakunakorn, and Maerkl 2019) and cell-like (Pautot, Frisken, and Weitz 2003) or subcellular mitochondria-like (Beneyton et al. 2018) droplets. These artificial compartments can be further designed to interact with each other via diffusively coupled (Tayar et al. 2017) or signaling-based (Adamala et al. 2017) mechanisms or to harbor key pieces of the prokaryotic division machinery (Godino et al. 2019). Besides direct encapsulation, these techniques can be used to generate droplets-in-droplets structures such that as many as 104 light-harvesting proteoliposomes can be encapsulated inside GUVs (Berhanu, Ueda, and Kuruma 2019). However, compartments produced in most studies have a fixed amount of membrane material. Bearing one ultimate goal of developing a synthetic cell capable of growth and division, it is also critical to create mechanisms for cells to form their own membranes. By engineering a soluble protein system, researchers have made feasible the de novo synthesis of phospholipid

21

membranes (Bhattacharya et al. 2019). Although this system does not require pre-existing membranes to function, it does need single-chain reactive lipid precursors, amine-functionalized lysolipids, and ATP. Moreover, the current implementation produces vesicles of various sizes and lamellarity in an uncontrolled way with the soluble enzyme Fad10 spontaneously associating with the de novo formed membranes. All these characteristics might pose difficulties for future implementations where controlled liposome composition and type is needed.

Cell size dominates the length scales of cellular life because any other cellular structures have to physically fit inside the cell. Understanding the relationship between subcellular structures and cell dimensions is crucial to build cells with functional subcellular compartments; on the other hand, synthetic cells also provide a uniquely powerful tool to investigate the mechanisms behind such a relationship. For example, the encapsulation of mitotic spindles in vesicles of different sizes showed that the spindle size scales with the compartment’s volume (Good et al. 2013). This points to the mechanisms in which cells sense and control their sizes. Some recent studies exploring this relationship suggested possible venues for implementing such mechanisms in synthetic cell systems. Cells utilize multiple strategies to control and maintain their size such as volume and area sensing (Facchetti et al. 2019), but also maintain specific ratios between subcellular structures and the total volume. A prominent example of the latter is the homeostasis of nuclear size relative to cytoplasmic volume (Cantwell and Nurse 2019a). By extensive screening of the genes involved in aberrant nuclear to cytoplasmic volume ratio, RNA processing and LINC complexes have been identified to regulate nuclear size. Additionally, different cytoplasmic factors are involved in nuclear size control. In particular, importin-alpha has been indicated as a key component in nuclear size homeostasis (Brownlee and Heald 2019). For a comprehensive review of the results in this area, see (Cantwell and Nurse 2019b). Finally, a key player in regulating cell size and shape is the cytoskeleton. A comprehensive review summarizing the recent efforts in cytoskeleton reconstitution in synthetic cells can be found here (Bashirzadeh and Liu 2019).

Besides micrometric scales, long-range communication and macroscopic pattern formation depend on microscopic rules between cell-sized agents. Therefore, by integrating cellular processes with agent communication, complex behaviors can be achieved such as those observed in biological communities. Moving forward, the homeostasis and size-control of synthetic cells and structures can only be achieved from a thorough understanding of their in vivo counterparts. The integration of processes across the different spatial scales of cellular life will ultimately lead to a complex functioning synthetic cell.

22

5.3 Perspectives on using mathematical modeling for bridging scales

Mathematical and computational modeling is key for studying artificial systems. Modeling can help in the design and explanation of synthetic constructs as well as making predictions that suggest new experiments. Additionally, a goal of synthetic biology is to produce standardized biological parts with robust and reproducible behaviors. Therefore, modeling is one of the main quantitative tools that facilitates the process of constructing synthetic cells.

Traditionally, the dynamics of chemical reactions and gene regulatory networks are described with ordinary differential equations (ODEs). In general, and when copy numbers are high, these models are useful. However, a recurrent problem is the parameterization of large sets of equations. Recently, new methods are used to solve this caveat, such as training a neural network with machine learning algorithms that can significantly improve computational efficiency (Wang et al. 2019). When dealing with small volumes, another branch of modeling becomes more relevant: stochastic modeling. Droplet compartmentalization can introduce partitioning noise that makes data interpretation difficult with ODE models (M. Weitz et al. 2014). In those cases, the stochastic analysis of experimental data becomes indispensable. Also, stochasticity sometimes produces results that are not present in mean-field ODE descriptions. For example, stochastic Turing patterns are studied and long-range order is found in regions of parameter space where a deterministic version of the model would not produce patterns (Karig et al. 2018).

When constructing a synthetic cell, the integration of spatially heterogeneous processes at different scales is a challenging task. The aforementioned tools typically work under assumptions of well-mixed reactions with single time/space discretization These gaps can be filled using two modeling approaches, namely, agent-based models (ABMs) and multiscale modeling. ABMs naturally include the heterogeneities of cellular processes while providing a framework for exploring agent-based rules. Importantly, large-scale behaviors and patterns can be studied as a function of cell-cell interaction rules. Multiscale modeling, instead, explicitly integrates simulations across the different temporal, spatial, and functional scales involved in a complex multiscale system. Using different levels of resolution in the same model, these techniques allow the observation of behaviors across multiple scales. For further reading on ABMs and multiscale modeling, please refer to (Gorochowski 2016) and (Walpole, Papin, and Peirce 2013), respectively. Capturing every aspect of cellular life in a single model appears unattainable. However, such an approach might be unnecessary as a model may only need to capture the important characteristics or a

23

phenomenon of interest. If heterogeneities (ABMs), noise (stochastic modeling), dimensions (multiscale approaches), and mean-field behaviors (ODEs) are considered when modeling, effective quantitative descriptions can be constructed to aid in the design of synthetic cell functions that accounts for different time and space scales.

6. FUTURE DIRECTIONS

In this review article, we described the engineering of synthetic cells as a natural progression of the study of biology by our increased ability to synthesize and manipulate biological components and integrating them into cell-like systems. The field is growing rapidly with creative approaches and new discoveries being generated. We summarize by highlighting some promising directions and also where challenges in synthetic cell research lie.

6.1 Challenges and issues to be addressed

1. A large portion of the research in synthetic biology focuses on genetic networks, leaving the realm of signaling systems less-explored; in particular, communication via post-translational modifications. Tools for engineering signaling systems to better control the relationship between reactions with fast and slow timescales are needed to improve the dynamic range and flexibility in the time domain of synthetic cells. 2. Protein modifications are one of the challenges with synthetizing signaling systems as compared to genetic systems. Advances in protein engineering and eukaryotic TX-TL systems will help in this direction. 3. Both 1 and 2 will help in the reconstitution of more complex eukaryotic systems such as an artificial cell cycle system from the bottom-up. 4. Integrating diverse modules remains a challenge because the cell-free conditions used to optimize the function of one module might not be suitable for another. 5. Optimizing TX-TL systems is needed for the production of controlled and coordinated expression of multiple protein operons or whole systems of operons at the same time. 6. Byproduct elimination and substrate/energy replenishment is a critical issue for creating synthetic cells with expanded life cycles and control thereof. 7. Functional expression of membrane proteins in synthetic cells is still challenging. 8. Robust control of self-dividing liposomes remains a challenge in the field. 9. Shape and size control in synthetic cells can be a fruitful challenge to understand both natural mechanisms and synthetic ones. 24

10. Methodologies for reliable scale-up of TX-TL systems are needed for all future endeavors.

6.2 Opportunities and areas to be further explored

Exciting new approaches towards protein-based compartmentalization and spatial control are on the horizon. In addition to the discovery and engineering of novel natural protein organelles or IDPs involved in LLPS, computational or de novo design of protein compartments and IDPs has recently gained attention (Koepnick et al. 2019; Baker 2019). Substantial advances have been made with regard to computationally designed protein shells and cages (Hsia et al. 2016; Bale et al. 2016; Brouwer et al. 2019; Cannon et al. 2020). In addition, novel design concepts based on symmetry relationships and metal-mediated assembly have also been achieved (Badieyan et al. 2017; Malay et al. 2019; Cannon, Ochoa, and Yeates 2019; Cristie-David et al. 2019; Cristie-David and Marsh 2019). So far, these systems have not been used in a synthetic cell context or as functional protein organelles with enzymatic cargo. The widespread adoption of de novo designed protein compartments will necessitate developing specific mechanisms with which to easily target arbitrary protein components to their interior. Another current limitation lies with the difficulty of designing protein shells larger than about 40 nm. IDPs represent even more challenging targets for computational protein design due to substantial limitations in our basic understanding of how IDP sequence is connected to structure, dynamics and ultimately function. However, some progress has been made with computationally designing IDPs (Schramm et al. 2017; Holehouse et al. 2017; Staller et al. 2018; Cohan, Ruff, and Pappu 2019). Future advances in computational protein design and engineering have the potential to address the issues raised above and yield more advanced and tailor-made protein components for controlling compartmentalization in cell- free systems and synthetic cells.

Although DNA is typically used as the information carrier, there are new ways to integrate DNA as part of synthetic cell architecture. Recently, a DNA-based cortex was incorporated into giant vesicles, endowing them with significantly higher mechanical stability towards osmotic shock (Kurokawa et al. 2017). Additionally, the powerful toolbox of DNA origami promises to enable the creation of structures supporting the membrane or even shaping the compartment (Dong et al. 2017; Noireaux and Liu 2020). The use of DNA as a structural material can extend our degree of control on the shape of synthetic cells and improve the flexibility in the range of stresses and stimuli they can withstand and respond to.

Many of the self-organizing principles required for a synthetic cell have been reconstituted to varying degrees. However, the buffer conditions, protein concentrations, and technical approaches (i.e.

25

microfluidic methodologies) have been specifically tailored towards one system and may therefore not be compatible with others. To produce a functional synthetic cell, the integration of processes across diverse scales is necessary. We believe multiscale modeling will provide important physical insights. Additionally, because high-order communication between synthetic cells is being explored, ABM becomes relevant. In the end, hybrid modeling approaches need to be employed to grasp the complexity of modularity and multi-scales that cellular life poses. Integration of different cellular processes will be a significant milestone achievement in synthetic cell research. Ideally, as an example, a successfully integrated reconstitution involves spatiotemporally coordinating the segregation of a sufficiently compacted minimal genome with the growth and division of a liposome.

7. CONCLUSION

The creation of synthetic cells is truly an interdisciplinary discipline built upon a strong foundation in physics, chemistry, and biology – but also closely relates to engineering. Besides the potential for deciphering design principles of natural living cells, synthetic cells have significant potential as valuable biomedical and biotechnological tools. Synthetic cells can be designed to be invisible to the immune system, compared to engineered living cells, or be decorated with specific antigens used for the development of vaccines. The versatility of synthetic cells is only limited by the imagination of their creators and this promises an exciting final phase of biology.

Funding Information

APL acknowledges funding from NSF (MCB-1935265; CBET-1844132; MCB-1817909; DMR-1939534). QY acknowledges funding from NSF (MCB-1817909; MCB-1553031), NIH (NIGMS- R35GM119688), and Alfred P. Sloan Foundation. AGV acknowledges funding from NSF (MCB-1817478 and CAREER Award MCB-1941966). TWG acknowledges funding from NIH (NIGMS-R35GM133325) and the Gordon and Betty Moore Foundation (Grant: 5506).

Acknowledgments

None.

26

Figures and Tables

Figure 1. Recent advances in constructing synthetic cells involve engineering synthetic cell membranes, development of compartmentalization strategies, and reconstituting dynamic processes in space and time. Black arrows depict the flow of molecules and the exchange of information happening across lipid/protein membranes or between the phase-separated liquid condensates of a membrane-less compartment.

27

Figure 2. Possible molecular composition for the boundaries of synthetic cells. Left. (a) A membrane composed by phospholipids and cholesterol is shown. Enlargements show the molecular structures of 1- palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine (POPC) and cholesterol. (b) Membrane structure of a proteinosome, showing the spatial organization of oleosin molecules. The enlargement shows a random peptide sequence. (c) Polymerosomal membrane composed by block copolymers. The enlargement shows the molecular structure of polystyrene-b-polyacrilic acid (PS-b-PAA). (d) A phospholipid/block copolymer hybrid membrane is shown. The illustrations are not drawn to scale. Right. Relative comparison of different characteristics between phospholipids, polypeptides, and block copolymers as a chassis material for synthetic cells.

Figure 3. Protein organelles and LLPS for designing compartmentalization in synthetic cells. A) The two known classes of protein organelles, encapsulin nanocompartments (Encs) and bacterial microcompartments (BMCs). B) Potential problems for enzymes and pathways arising without compartmentalization. C) The advantages of compartmentalizing metabolic pathways. D) Integration of different types of designed and modular protein-based compartmentalization systems for optimizing synthetic cell metabolism.

28

Figure 4. Examples of Min/FtsZ system reconstitutions. A) MinD and MinE have been shown to form a variety of patterns when reconstituted in cell-free setups. In this example, GFP-MinD (cyan) and MinE- Alexa647 (magenta) were pre-incubated with ATP and infused into a flowcell coated with a supported lipid bilayer (SLB). Still images show the different types of patterns supported by the decreasing protein density in the flowcell from inlet to outlet. Adapted with permission from (Mizuuchi and Vecchiarelli 2018). B) Reconstitution of Min oscillations in a rod-shaped trough has been shown to confine FtsZ-YFP- MTS polymerization to the middle of the compartment. Adapted from (Zieske and Schwille 2014) under the CC BY 4.0 license.

29

Figure 5. Several temporal scales involved in biological cells (left) and synthetic cells (right). The in vitro reconstituted phenomena range from oscillations with multi-second periods up to multicellular structures whose aggregation unfolds in the order of days. In between, temporal scales range from metabolic processes that can be sustained for several hours, the oscillations of coupled micro- compartments, the formation of self-organized structures in Xenopus laevis egg extracts, and the fast timescale response of the engineered phosphorylation regulatory networks. Some illustration images in were created with BioRender.com.

30

Figure 6. A range of spatial scales in natural cellular life (left) and in synthetic cells (right). In the synthetic cell context, nanometric functional compartments have been generated as well as millimeter- sized protein expression patterns. Energy modules were constructed with nanometer-sized proteoliposomes and inverted membranes. Synthetic cells or Synells have also been developed which can interact with each other. Additionally, a plausible mechanism for de novo liposome production was developed yielding compartments of compatible size with biological cells. Subcellular structures and their dependence on cell volume were studied by reconstructing mitotic spindles in droplets. Finally, protein expression patterns, compartments capable of mitosis, and multicellular structures have been produced in relevant spatial scales where eukaryotic cells and tissues operate. Some illustration images in were created with BioRender.com.

31

References

Adamala, Katarzyna P., Daniel A. Martin-Alarcon, Katriona R. Guthrie-Honea, and Edward S. Boyden. 2017. “Engineering Genetic Circuit Interactions within and between Synthetic Minimal Cells.” Nature Chemistry 9 (5): 431–39. https://doi.org/10.1038/nchem.2644.

Ahmad, Mohd Ridzuan, Masahiro Nakajima, Seiji Kojima, Michio Homma, and Toshio Fukuda. 2008. “The Effects of Cell Sizes, Environmental Conditions, and Growth Phases on the Strength of Individual W303 Yeast Cells inside ESEM.” IEEE Transactions on Nanobioscience 7 (3): 185–93. https://doi.org/10.1109/TNB.2008.2002281.

Akita, Fusamichi, Khoon Tee Chong, Hideaki Tanaka, Eiki Yamashita, Naoyuki Miyazaki, Yuichiro Nakaishi, Mamoru Suzuki, et al. 2007. “The Crystal Structure of a Virus-like Particle from the Hyperthermophilic Archaeon Pyrococcus Furiosus Provides Insight into the Evolution of Viruses.” Journal of Molecular Biology 368 (5): 1469–83. https://doi.org/10.1016/j.jmb.2007.02.075.

Amit, Roee, Amos B. Oppenheim, and Joel Stavans. 2003. “Increased Bending Rigidity of Single DNA Molecules by H-NS, a Temperature and Osmolarity Sensor.” Biophysical Journal 84 (4): 2467–73. https://doi.org/10.1016/S0006-3495(03)75051-6.

Aussignargues, Cléement, Bradley C. Paasch, Raul Gonzalez-Esquer, Onur Erbilgin, and Cheryl A. Kerfeld. 2015. “Bacterial Microcompartment Assembly: The Key Role of Encapsulation Peptides.” Communicative and Integrative Biology 8 (3). https://doi.org/10.1080/19420889.2015.1039755.

Aussignargues, Clément, Maria-Eirini Pandelia, Markus Sutter, Jefferson S Plegaria, Jan Zarzycki, Aiko Turmo, Jingcheng Huang, et al. 2016. “Structure and Function of a Bacterial Microcompartment Shell Protein Engineered to Bind a [4Fe-4S] Cluster.” Journal of the American Chemical Society 138 (16): 5262– 70. https://doi.org/10.1021/jacs.5b11734.

Axen, Seth D., Onur Erbilgin, and Cheryl A. Kerfeld. 2014. “A Taxonomy of Bacterial Microcompartment Loci Constructed by a Novel Scoring Method.” PLoS Computational Biology 10 (10): e1003898. https://doi.org/10.1371/journal.pcbi.1003898.

Badieyan, Somayesadat, Aaron Sciore, Joseph D. Eschweiler, Philipp Koldewey, Ajitha S. Cristie-David, Brandon T. Ruotolo, James C.A. Bardwell, Min Su, and E. Neil G. Marsh. 2017. “Symmetry-Directed Self- Assembly of a Tetrahedral Protein Cage Mediated by de Novo-Designed Coiled Coils.” ChemBioChem 18 (19): 1888–92. https://doi.org/10.1002/cbic.201700406.

32

Bae, Yoonji, Gwang Joong Kim, Hansol Kim, Seong Guk Park, Hyun Suk Jung, and Sebyung Kang. 2018. “Engineering Tunable Dual Functional Protein Cage Nanoparticles Using Bacterial Superglue.” Biomacromolecules 19 (7): 2896–2904. https://doi.org/10.1021/acs.biomac.8b00457.

Baker, David. 2019. “What Has de Novo Protein Design Taught Us about Protein Folding and Biophysics?” Protein Science : A Publication of the Protein Society 28 (4): 678–83. https://doi.org/10.1002/pro.3588.

Bale, Jacob B., Shane Gonen, Yuxi Liu, William Sheffler, Daniel Ellis, Chantz Thomas, Duilio Cascio, et al. 2016. “Accurate Design of Megadalton-Scale Two-Component Icosahedral Protein Complexes.” Science 353 (6297): 389–94. https://doi.org/10.1126/science.aaf8818.

Bashirzadeh, Yashar, and Allen P. Liu. 2019. “Encapsulation of the Cytoskeleton: Towards Mimicking the Mechanics of a Cell.” Soft Matter 15 (42): 8425–36. https://doi.org/10.1039/c9sm01669d.

Baxter, Jamie C., and Barbara E. Funnell. 2014. “Plasmid Partition Mechanisms.” Microbiology Spectrum 2 (6). https://doi.org/10.1128/microbiolspec.plas-0023-2014.

Bellomo, Enrico G., Michael D. Wyrsta, Lisa Pakstis, Darrin J. Pochan, and Timothy J. Deming. 2004. “Stimuli-Responsive Polypeptide Vesicles by Conformation-Specific Assembly.” Nature Materials 3 (4): 244–48. https://doi.org/10.1038/nmat1093.

Beneyton, Thomas, Dorothee Krafft, Claudia Bednarz, Christin Kleineberg, Christian Woelfer, Ivan Ivanov, Tanja Vidaković-Koch, Kai Sundmacher, and Jean Christophe Baret. 2018. “Out-of-Equilibrium Microcompartments for the Bottom-up Integration of Metabolic Functions.” Nature Communications 9 (1): 1–10. https://doi.org/10.1038/s41467-018-04825-1.

Berhanu, Samuel, Takuya Ueda, and Yutetsu Kuruma. 2019. “Artificial Photosynthetic Cell Producing Energy for Protein Synthesis.” Nature Communications 10 (1): 1–10. https://doi.org/10.1038/s41467- 019-09147-4.

Bhardwaj, Tulika, and Pallavi Somvanshi. 2017. “Pan-Genome Analysis of Clostridium Botulinum Reveals Unique Targets for Drug Development.” Gene 623 (August): 48–62. https://doi.org/10.1016/j.gene.2017.04.019.

Bhattacharya, Ahanjit, Roberto J. Brea, Henrike Niederholtmeyer, and Neal K. Devaraj. 2019. “A Minimal Biochemical Route towards de Novo Formation of Synthetic Phospholipid Membranes.” Nature Communications 10 (1): 1–8. https://doi.org/10.1038/s41467-018-08174-x. 33

Blaauwen, Tanneke den, Leendert W. Hamoen, and Petra Anne Levin. 2017. “The Divisome at 25: The Road Ahead.” Current Opinion in Microbiology. Elsevier Ltd. https://doi.org/10.1016/j.mib.2017.01.007.

Blaauwen, Tanneke den, and Joen Luirink. 2019. “Checks and Balances in Bacterial Cell Division.” MBio 10 (1). https://doi.org/10.1128/mBio.00149-19.

Bobik, Thomas A, Brent P Lehman, and Todd O Yeates. 2015. “Bacterial Microcompartments: Widespread Prokaryotic Organelles for Isolation and Optimization of Metabolic Pathways.” Molecular Microbiology 98 (2): 193–207. https://doi.org/10.1111/mmi.13117.

Brangwynne, Clifford P., Christian R. Eckmann, David S. Courson, Agata Rybarska, Carsten Hoege, Jöbin Gharakhani, Frank Jülicher, and Anthony A. Hyman. 2009. “Germline P Granules Are Liquid Droplets That Localize by Controlled Dissolution/Condensation.” Science 324 (5935): 1729–32. https://doi.org/10.1126/science.1172046.

Brouwer, Philip J.M., Aleksandar Antanasijevic, Zachary Berndsen, Anila Yasmeen, Brooke Fiala, Tom P.L. Bijl, Ilja Bontjer, et al. 2019. “Enhancing and Shaping the Immunogenicity of Native-like HIV-1 Envelope Trimers with a Two-Component Protein Nanoparticle.” Nature Communications 10 (1). https://doi.org/10.1038/s41467-019-12080-1.

Brownlee, Christopher, and Rebecca Heald. 2019. “Importin α Partitioning to the Plasma Membrane Regulates Intracellular Scaling.” Cell 176 (4): 805-815.e8. https://doi.org/10.1016/j.cell.2018.12.001.

Cai, Fei, Susan L Bernstein, Steven C Wilson, and Cheryl A Kerfeld. 2016. “Production and Characterization of Synthetic Carboxysome Shells with Incorporated Luminal Proteins.” Plant Physiology 170 (3): 1868–77. https://doi.org/10.1104/pp.15.01822.

Cai, Fei, Markus Sutter, Susan L. Bernstein, James N. Kinney, and Cheryl A. Kerfeld. 2015. “Engineering Bacterial Microcompartment Shells: Chimeric Shell Proteins and Chimeric Carboxysome Shells.” ACS Synthetic Biology 4 (4): 444–53. https://doi.org/10.1021/sb500226j.

Cannon, Kevin A., Rachel U. Park, Scott E. Boyken, Una Nattermann, Sue Yi, David Baker, Neil P. King, and Todd O. Yeates. 2020. “Design and Structure of Two New Protein Cages Illustrate Successes and Ongoing Challenges in Protein Engineering.” Protein Science 29 (4): 919–29. https://doi.org/10.1002/pro.3802.

Cannon, Kevin A, Jessica M Ochoa, and Todd O Yeates. 2019. “High-Symmetry Protein Assemblies: Patterns and Emerging Applications.” Current Opinion in Structural Biology 55 (April): 77–84. https://doi.org/10.1016/j.sbi.2019.03.008. 34

Cantwell, Helena, and Paul Nurse. 2019a. “A Systematic Genetic Screen Identifies Essential Factors Involved in Nuclear Size Control.” Edited by Jan M. Skotheim. PLOS Genetics 15 (2): e1007929. https://doi.org/10.1371/journal.pgen.1007929.

Cantwell, Helena, and Paul Nurse. 2019b. “Unravelling Nuclear Size Control.” Current Genetics. Springer Verlag. https://doi.org/10.1007/s00294-019-00999-3.

Caspi, Yaron, and Cees Dekker. 2016. “Mapping out Min Protein Patterns in Fully Confined Fluidic Chambers.” ELife 5 (NOVEMBER2016). https://doi.org/10.7554/eLife.19271.

Cassidy-Amstutz, Caleb, Luke Oltrogge, Catherine C Going, Antony Lee, Poh Teng, David Quintanilla, Alexandra East-Seletsky, Evan R Williams, and David F Savage. 2016. “Identification of a Minimal Peptide Tag for in Vivo and in Vitro Loading of Encapsulin.” Biochemistry 55 (24): 3461–68. https://doi.org/10.1021/acs.biochem.6b00294.

Chen, Cheng, Joshua S MacCready, Daniel C Ducat, and Katherine W Osteryoung. 2018. “The Molecular Machinery of Chloroplast Division.” Plant Physiology 176 (1): 138–51. https://doi.org/10.1104/pp.17.01272.

Chen, Irene A., Kourosh Salehi-Ashtiani, and Jack W. Szostak. 2005. “RNA Catalysis in Model Protocell Vesicles.” Journal of the American Chemical Society 127 (38): 13213–19. https://doi.org/10.1021/ja051784p.

Chen, Zhaowei, Jinqiang Wang, Wujin Sun, Edikan Archibong, Anna R. Kahkoska, Xudong Zhang, Yue Lu, Frances S. Ligler, John B. Buse, and Zhen Gu. 2018. “Synthetic Beta Cells for Fusion-Mediated Dynamic Insulin Secretion.” Nature Chemical Biology 14 (1): 86–93. https://doi.org/10.1038/nchembio.2511.

Cheng, Xianrui, and James E. Ferrell. 2019. “Spontaneous Emergence of Cell-like Organization in Xenopus Egg Extracts.” Science 366 (6465): 631–37. https://doi.org/10.1126/science.aav7793.

Choi, Bongseo, Hansol Kim, Hyukjun Choi, and Sebyung Kang. 2018. “Protein Cage Nanoparticles as Delivery Nanoplatforms.” In Advances in Experimental Medicine and Biology, 1064:27–43. Springer New York LLC. https://doi.org/10.1007/978-981-13-0445-3_2.

Choi, Bongseo, Hyojin Moon, Sung Joon Hong, Changsik Shin, Yoonkyung Do, Seongho Ryu, and Sebyung Kang. 2016. “Effective Delivery of Antigen-Encapsulin Nanoparticle Fusions to Dendritic Cells Leads to Antigen-Specific Cytotoxic T Cell Activation and Tumor Rejection.” ACS Nano 10 (8): 7339–50. https://doi.org/10.1021/acsnano.5b08084. 35

Choi, Hyo Jick, and Carlo D. Montemagno. 2007. “Light-Driven Hybrid Bioreactor Based on Protein- Incorporated Polymer Vesicles.” IEEE Transactions on 6 (2): 171–76. https://doi.org/10.1109/TNANO.2007.891822.

Choudhary, Swati, Maureen B Quin, Mark A Sanders, Ethan T Johnson, and Claudia Schmidt-Dannert. 2012. “Engineered Protein Nano-Compartments for Targeted Enzyme Localization.” PloS One 7 (3): e33342. https://doi.org/10.1371/journal.pone.0033342.

Chowdhury, C., S. Sinha, S. Chun, T. O. Yeates, and T. A. Bobik. 2014. “Diverse Bacterial Microcompartment Organelles.” Microbiology and Molecular Biology Reviews 78 (3): 438–68. https://doi.org/10.1128/mmbr.00009-14.

Cohan, Megan C, Kiersten M Ruff, and Rohit V Pappu. 2019. “Information Theoretic Measures for Quantifying Sequence-Ensemble Relationships of Intrinsically Disordered Proteins.” Protein Engineering, Design & Selection : PEDS 32 (4): 191–202. https://doi.org/10.1093/protein/gzz014.

Contreras, Heidi, Matthew S Joens, Lisa M McMath, Vincent P Le, Michael V Tullius, Jaqueline M Kimmey, Neda Bionghi, Marcus A Horwitz, James A J Fitzpatrick, and Celia W Goulding. 2014. “Characterization of a Mycobacterium Tuberculosis Nanocompartment and Its Potential Cargo Proteins.” The Journal of Biological Chemistry 289 (26): 18279–89. https://doi.org/10.1074/jbc.M114.570119.

Cornejo, Elias, Nicole Abreu, and Arash Komeili. 2014. “Compartmentalization and Organelle Formation in Bacteria.” Current Opinion in Cell Biology 26 (1): 132–38. https://doi.org/10.1016/j.ceb.2013.12.007.

Cristie-David, Ajitha S, Junjie Chen, Derek B Nowak, Amy L Bondy, Kai Sun, Sung I Park, Mark M Banaszak Holl, Min Su, and E Neil G Marsh. 2019. “Coiled-Coil-Mediated Assembly of an Icosahedral Protein Cage with Extremely High Thermal and Chemical Stability.” Journal of the American Chemical Society 141 (23): 9207–16. https://doi.org/10.1021/jacs.8b13604.

Cristie-David, Ajitha S, and E Neil G Marsh. 2019. “Metal-Dependent Assembly of a Protein Nano-Cage.” Protein Science : A Publication of the Protein Society 28 (9): 1620–29. https://doi.org/10.1002/pro.3676.

Crowe, Charles D, and Christine D Keating. 2018. “Liquid-Liquid Phase Separation in Artificial Cells.” Interface Focus 8 (5): 20180032. https://doi.org/10.1098/rsfs.2018.0032.

Cunha, Sónia, Conrad L. Woldringh, and Theo Odijk. 2001. “Polymer-Mediated Compaction and Internal Dynamics of Isolated Escherichia Coli Nucleoids.” Journal of Structural Biology 136 (1): 53–66. 36

https://doi.org/10.1006/jsbi.2001.4420.

Dame, Remus T., Maarten C. Noom, and Gijs J.L. Wuite. 2006. “Bacterial Chromatin Organization by H- NS Protein Unravelled Using Dual DNA Manipulation.” Nature 444 (7117): 387–90. https://doi.org/10.1038/nature05283.

Danielli, J.F. 1972. “The Artificial Synthesis of New Life Forms In Relation to Social and Industrial Evolution.” In The Future of Man, 95–112. Elsevier. https://doi.org/10.1016/b978-0-12-229060-2.50012- 4.

Deng, Nan-Nan, and Wilhelm T S Huck. 2017. “Microfluidic Formation of Monodisperse Coacervate Organelles in Liposomes.” Angewandte Chemie (International Ed. in English) 56 (33): 9736–40. https://doi.org/10.1002/anie.201703145.

Deshpande, Siddharth, Frank Brandenburg, Anson Lau, Mart G F Last, Willem Kasper Spoelstra, Louis Reese, Sreekar Wunnava, Marileen Dogterom, and Cees Dekker. 2019. “Spatiotemporal Control of Coacervate Formation within Liposomes.” Nature Communications 10 (1): 1800. https://doi.org/10.1038/s41467-019-09855-x.

Diekmann, Yoan, and José B Pereira-Leal. 2013. “Evolution of Intracellular Compartmentalization.” The Biochemical Journal 449 (2): 319–31. https://doi.org/10.1042/BJ20120957.

Dong, Yuanchen, Yuhe Renee Yang, Yiyang Zhang, Dianming Wang, Xixi Wei, Saswata Banerjee, Yan Liu, Zhongqiang Yang, Hao Yan, and Dongsheng Liu. 2017. “Cuboid Vesicles Formed by Frame-Guided Assembly on DNA Origami Scaffolds.” Angewandte Chemie International Edition 56 (6): 1586–89. https://doi.org/10.1002/anie.201610133.

Dryden, Kelly A, Christopher S Crowley, Shiho Tanaka, Todd O Yeates, and Mark Yeager. 2009. “Two- Dimensional Crystals of Carboxysome Shell Proteins Recapitulate the Hexagonal Packing of Three- Dimensional Crystals.” Protein Science : A Publication of the Protein Society 18 (12): 2629–35. https://doi.org/10.1002/pro.272.

Du, Mingjian, and Zhijian J. Chen. 2018. “DNA-Induced Liquid Phase Condensation of CGAS Activates Innate Immune Signaling.” Science 361 (6403): 704–9. https://doi.org/10.1126/science.aat1022.

Dwidar, Mohammed, Yusuke Seike, Shungo Kobori, Charles Whitaker, Tomoaki Matsuura, and Yohei Yokobayashi. 2019. “Programmable Artificial Cells Using Histamine-Responsive Synthetic Riboswitch.” Journal of the American Chemical Society 141 (28): 11103–14. https://doi.org/10.1021/jacs.9b03300. 37

Einfalt, T., D. Witzigmann, C. Edlinger, S. Sieber, R. Goers, A. Najer, M. Spulber, O. Onaca-Fischer, J. Huwyler, and C. G. Palivan. 2018. “Biomimetic Artificial Organelles with in Vitro and in Vivo Activity Triggered by Reduction in Microenvironment.” Nature Communications 9 (1): 1–12. https://doi.org/10.1038/s41467-018-03560-x.

Einfalt, Tomaž, Roland Goers, Ionel Adrian Dinu, Adrian Najer, Mariana Spulber, Ozana Onaca-Fischer, and Cornelia G. Palivan. 2015. “Stimuli-Triggered Activity of Nanoreactors by Biomimetic Engineering Polymer Membranes.” Nano Letters 15 (11): 7596–7603. https://doi.org/10.1021/acs.nanolett.5b03386.

Elbaum-Garfinkle, Shana, Younghoon Kim, Krzysztof Szczepaniak, Carlos Chih-Hsiung Chen, Christian R Eckmann, Sua Myong, and Clifford P Brangwynne. 2015. “The Disordered P Granule Protein LAF-1 Drives Phase Separation into Droplets with Tunable Viscosity and Dynamics.” Proceedings of the National Academy of Sciences of the United States of America 112 (23): 7189–94. https://doi.org/10.1073/pnas.1504822112.

Elston, Timothy C., and George Oster. 1997. “Protein Turbines I: The Bacterial Flagellar Motor.” Biophysical Journal 73 (2): 703–21. https://doi.org/10.1016/S0006-3495(97)78104-9.

Erb, Marcella L., James A. Kraemer, Joanna K.C. Coker, Vorrapon Chaikeeratisak, Poochit Nonejuie, David A. Agard, and Joe Pogliano. 2014. “A Bacteriophage Tubulin Harnesses Dynamic Instability to Center DNA in Infected Cells.” ELife 3. https://doi.org/10.7554/eLife.03197.

Facchetti, Giuseppe, Benjamin Knapp, Ignacio Flor-Parra, Fred Chang, and Martin Howard. 2019. “Reprogramming Cdr2-Dependent Geometry-Based Cell Size Control in Fission Yeast.” Current Biology 29 (2): 350-358.e4. https://doi.org/10.1016/j.cub.2018.12.017.

Fang, Yi, Fang Huang, Matthew Faulkner, Qiuyao Jiang, Gregory F Dykes, Mengru Yang, and Lu-Ning Liu. 2018. “Engineering and Modulating Functional Cyanobacterial CO2-Fixing Organelles.” Frontiers in Plant Science 9 (June): 739. https://doi.org/10.3389/fpls.2018.00739.

Fink, Gero, and Jan Löwe. 2015. “Reconstitution of a Prokaryotic Minus End-Tracking System Using TubRC Centromeric Complexes and Tubulin-like Protein TubZ Filaments.” Proceedings of the National Academy of Sciences of the United States of America 112 (15): E1845-50. https://doi.org/10.1073/pnas.1423746112.

Gall, Antoine Le, Diego I Cattoni, Baptiste Guilhas, Céline Mathieu-Demazière, Laura Oudjedi, Jean- Bernard Fiche, Jérôme Rech, et al. 2016. “Bacterial Partition Complexes Segregate within the Volume of

38

the Nucleoid.” Nature Communications 7 (July): 12107. https://doi.org/10.1038/ncomms12107.

Ganji, Mahipal, Indra A. Shaltiel, Shveta Bisht, Eugene Kim, Ana Kalichava, Christian H. Haering, and Cees Dekker. 2018. “Real-Time Imaging of DNA Loop Extrusion by Condensin.” Science 360 (6384): 102–5. https://doi.org/10.1126/science.aar7831.

Garamella, Jonathan, Sagardip Majumder, Allen P. Liu, and Vincent Noireaux. 2019. “An Adaptive Synthetic Cell Based on Mechanosensing, Biosensing, and Inducible Gene Circuits.” ACS Synthetic Biology 8 (8): 1913–20. https://doi.org/10.1021/acssynbio.9b00204.

Garamella, Jonathan, Ryan Marshall, Mark Rustad, and Vincent Noireaux. 2016. “The All E. Coli TX-TL Toolbox 2.0: A Platform for Cell-Free Synthetic Biology.” ACS Synthetic Biology 5 (4): 344–55. https://doi.org/10.1021/acssynbio.5b00296.

Garner, Ethan C., Christopher S Campbell, Douglas B Weibel, and R Dyche Mullins. 2007. “Reconstitution of DNA Segregation.” Science 315 (March): 1270–74. https://doi.org/10.1126/science.1138527.

Garner, Ethan C, Christopher S Campbell, and R Dyche Mullins. 2004. “Dynamic Instability in a DNA- Segregating Prokaryotic Actin Homolog.” Science (New York, N.Y.) 306 (5698): 1021–25. https://doi.org/10.1126/science.1101313.

Gerdes, Kenn, Martin Howard, and Florian Szardenings. 2010. “Pushing and Pulling in Prokaryotic DNA Segregation.” Cell. https://doi.org/10.1016/j.cell.2010.05.033.

Giessen, Tobias W., and Pamela A. Silver. 2016a. “Encapsulation as a Strategy for the Design of Biological Compartmentalization.” Journal of Molecular Biology. Academic Press. https://doi.org/10.1016/j.jmb.2015.09.009.

Giessen, Tobias W. 2016. “Encapsulins: Microbial Nanocompartments with Applications in Biomedicine, Nanobiotechnology and Materials Science.” Current Opinion in Chemical Biology 34 (October): 1–10. https://doi.org/10.1016/j.cbpa.2016.05.013.

Giessen, Tobias W, and Pamela A Silver. 2016b. “Converting a Natural Protein Compartment into a Nanofactory for the Size-Constrained Synthesis of Antimicrobial Silver Nanoparticles.” ACS Synthetic Biology 5 (12): 1497–1504. https://doi.org/10.1021/acssynbio.6b00117.

Giessen, Tobias W, and Pamela A Silver. 2017a. “Widespread Distribution of Encapsulin Nanocompartments Reveals Functional Diversity.” Nature Microbiology 2 (March): 17029.

39

https://doi.org/10.1038/nmicrobiol.2017.29.

Giessen, Tobias W, and Pamela A Silver. 2017b. “Engineering Carbon Fixation with Artificial Protein Organelles.” Current Opinion in Biotechnology 46 (August): 42–50. https://doi.org/10.1016/j.copbio.2017.01.004.

Gligoris, Thomas G, Johanna C Scheinost, Frank Bürmann, Naomi Petela, Kok-Lung Chan, Pelin Uluocak, Frédéric Beckouët, Stephan Gruber, Kim Nasmyth, and Jan Löwe. 2014. “Closing the Cohesin Ring: Structure and Function of Its Smc3-Kleisin Interface.” Science (New York, N.Y.) 346 (6212): 963–67. https://doi.org/10.1126/science.1256917.

Godino, Elisa, Jonás Noguera López, David Foschepoth, Céline Cleij, Anne Doerr, Clara Ferrer Castellà, and Christophe Danelon. 2019. “De Novo Synthesized Min Proteins Drive Oscillatory Liposome Deformation and Regulate FtsA-FtsZ Cytoskeletal Patterns.” Nature Communications 10 (1): 1–12. https://doi.org/10.1038/s41467-019-12932-w.

Godino, Elisa, Jonás Noguera López, Ilias Zarguit, Anne Doerr, Mercedes Jimenez, Germán Rivas, and Christophe Danelon. 2020. “Cell-Free Biogenesis of Bacterial Division Proto-Rings That Can Constrict Liposomes.” Communications Biology 3 (1): 539. https://doi.org/10.1038/s42003-020-01258-9.

Good, Matthew C., Michael D. Vahey, Arunan Skandarajah, Daniel A. Fletcher, and Rebecca Heald. 2013. “Cytoplasmic Volume Modulates Spindle Size during Embryogenesis.” Science 342 (6160): 856–60. https://doi.org/10.1126/science.1243147.

Göpfrich, Kerstin, Ilia Platzman, and Joachim P. Spatz. 2018. “Mastering Complexity: Towards Bottom-up Construction of Multifunctional Eukaryotic Synthetic Cells.” Trends in Biotechnology. Elsevier Ltd. https://doi.org/10.1016/j.tibtech.2018.03.008.

Gordley, Russell M., Reid E. Williams, Caleb J. Bashor, Jared E. Toettcher, Shude Yan, and Wendell A. Lim. 2016. “Engineering Dynamical Control of Cell Fate Switching Using Synthetic Phospho-Regulons.” Proceedings of the National Academy of Sciences of the United States of America 113 (47): 13528–33. https://doi.org/10.1073/pnas.1610973113.

Gorochowski, Thomas E. 2016. “Agent-Based Modelling in Synthetic Biology.” Essays in Biochemistry 60 (4): 325–36. https://doi.org/10.1042/EBC20160037.

Guan, Ye, Zhengda Li, Shiyuan Wang, Patrick M. Barnes, Xuwen Liu, Haotian Xu, Minjun Jin, Allen P. Liu, and Qiong Yang. 2018. “A Robust and Tunable Mitotic Oscillator in Artificial Cells.” ELife 7 (April). 40

https://doi.org/10.7554/eLife.33549.

Guan, Ye, Shiyuan Wang, Minjun Jin, Haotian Xu, and Qiong Yang. 2018. “Reconstitution of Cell-Cycle Oscillations in Microemulsions of Cell-Free Xenopus Egg Extracts.” Journal of Visualized Experiments 2018 (139). https://doi.org/10.3791/58240.

Guet, Călin C, Luke Bruneaux, Taejin L Min, Dan Siegal-Gaskins, Israel Figueroa, Thierry Emonet, and Philippe Cluzel. 2008. “Minimally Invasive Determination of MRNA Concentration in Single Living Bacteria.” Nucleic Acids Research 36 (12): e73. https://doi.org/10.1093/nar/gkn329.

Gulliver, G. 1875. “On the Size and Shape of Red Corpuscles of the Blood of Vertebrates, with Drawings of Them to a Uniform Scale, and Extended and Revised Tables of Measurements.” Proceedings of the Zoological Society of London 1875: 474–95.

Hagen, Andrew, Jefferson S Plegaria, Nancy Sloan, Bryan Ferlez, Clement Aussignargues, Rodney Burton, and Cheryl A Kerfeld. 2018. “In Vitro Assembly of Diverse Bacterial Microcompartment Shell Architectures.” Nano Letters 18 (11): 7030–37. https://doi.org/10.1021/acs.nanolett.8b02991.

Hagen, Andrew, Markus Sutter, Nancy Sloan, and Cheryl A Kerfeld. 2018. “Programmed Loading and Rapid Purification of Engineered Bacterial Microcompartment Shells.” Nature Communications 9 (1): 2881. https://doi.org/10.1038/s41467-018-05162-z.

Harvey, P C, M Watson, S Hulme, M A Jones, M Lovell, A Berchieri, J Young, N Bumstead, and P Barrow. 2011. “Salmonella Enterica Serovar Typhimurium Colonizing the Lumen of the Chicken Intestine Grows Slowly and Upregulates a Unique Set of Virulence and Metabolism Genes.” Infection and Immunity 79 (10): 4105–21. https://doi.org/10.1128/IAI.01390-10.

Heldt, Dana, Stefanie Frank, Arefeh Seyedarabi, Dimitrios Ladikis, Joshua B Parsons, Martin J Warren, and Richard W Pickersgill. 2009. “Structure of a Trimeric Bacterial Microcompartme1. Heldt D, Frank S, Se1. Heldt D, Frank S, Seyedarabi A, et Al. Structure of a Trimeric Bacterial Microcompartme1. Heldt D, Frank S, Seyedarabi A, et Al. Structure of a Trimeric Bacterial Microcompartment S.” The Biochemical Journal 423 (2): 199–207. https://doi.org/10.1042/BJ20090780.

Herring, Taylor I., Tiffany N. Harris, Chiranjit Chowdhury, Sujit Kumar Mohanty, and Thomas A. Bobik. 2018. “A Bacterial Microcompartment Is Used for Choline Fermentation by Escherichia Coli 536.” Journal of Bacteriology 200 (10). https://doi.org/10.1128/JB.00764-17.

Hilburger, Claire E., Miranda L. Jacobs, Kamryn R. Lewis, Justin A. Peruzzi, and Neha P. Kamat. 2019. 41

“Controlling Secretion in Artificial Cells with a Membrane and Gate.” ACS Synthetic Biology 8 (6): 1224– 30. https://doi.org/10.1021/acssynbio.8b00435.

Hindley, James W., Daniela G. Zheleva, Yuval Elani, Kalypso Charalambous, Laura M.C. Barter, Paula J. Booth, Charlotte L. Bevan, Robert V. Law, and Oscar Ces. 2019. “Building a Synthetic Mechanosensitive Signaling Pathway in Compartmentalized Artificial Cells.” Proceedings of the National Academy of Sciences of the United States of America 116 (34): 16711–16. https://doi.org/10.1073/pnas.1903500116.

Hirsch, Nicolas, Lyle B. Zimmerman, and Robert M. Grainger. 2002. “Xenopus, the next Generation: X. Tropicalis Genetics and Genomics.” Developmental Dynamics. https://doi.org/10.1002/dvdy.10178.

Holehouse, Alex S, Rahul K Das, James N Ahad, Mary O G Richardson, and Rohit V Pappu. 2017. “CIDER: Resources to Analyze Sequence-Ensemble Relationships of Intrinsically Disordered Proteins.” Biophysical Journal 112 (1): 16–21. https://doi.org/10.1016/j.bpj.2016.11.3200.

Hsia, Yang, Jacob B Bale, Shane Gonen, Dan Shi, William Sheffler, Kimberly K Fong, Una Nattermann, et al. 2016. “Design of a Hyperstable 60-Subunit Protein Dodecahedron. [Corrected].” Nature 535 (7610): 136–39. https://doi.org/10.1038/nature18010.

Hu, Zonglin, Edward P Gogol, and Joe Lutkenhaus. 2002. “Dynamic Assembly of MinD on Phospholipid Vesicles Regulated by ATP and MinE.” Proceedings of the National Academy of Sciences of the United States of America 99 (10): 6761–66. https://doi.org/10.1073/pnas.102059099.

Hu, Zonglin, and Joe Lutkenhaus. 2003. “A Conserved Sequence at the C-Terminus of MinD Is Required for Binding to the Membrane and Targeting MinC to the Septum.” Molecular Microbiology 47 (2): 345– 55. https://doi.org/10.1046/j.1365-2958.2003.03321.x.

Hutchison, Clyde A., Ray Yuan Chuang, Vladimir N. Noskov, Nacyra Assad-Garcia, Thomas J. Deerinck, Mark H. Ellisman, John Gill, et al. 2016. “Design and Synthesis of a Minimal Bacterial Genome.” Science 351 (6280). https://doi.org/10.1126/science.aad6253.

Hwang, Ling Chin, Anthony G Vecchiarelli, Yong-Woon Han, Michiyo Mizuuchi, Yoshie Harada, Barbara E Funnell, and Kiyoshi Mizuuchi. 2013. “ParA-Mediated Plasmid Partition Driven by Protein Pattern Self- Organization.” The EMBO Journal 32 (9): 1238–49. https://doi.org/10.1038/emboj.2013.34.

Hyman, Anthony A, Christoph A Weber, and Frank Jülicher. 2014. “Liquid-Liquid Phase Separation in Biology.” Annual Review of Cell and 30 (1): 39–58. https://doi.org/10.1146/annurev-cellbio-100913-013325. 42

Iancu, Cristina V, Dylan M Morris, Zhicheng Dou, Sabine Heinhorst, Gordon C Cannon, and Grant J Jensen. 2010. “Organization, Structure, and Assembly of Alpha-Carboxysomes Determined by Electron Cryotomography of Intact Cells.” Journal of Molecular Biology 396 (1): 105–17. https://doi.org/10.1016/j.jmb.2009.11.019.

Jacobs, Miranda L., Margrethe A. Boyd, and Neha P. Kamat. 2019. “Diblock Copolymers Enhance Folding of a Mechanosensitive Membrane Protein during Cell-Free Expression.” Proceedings of the National Academy of Sciences of the United States of America 116 (10): 4031–36. https://doi.org/10.1073/pnas.1814775116.

Jain, Ankur, and Ronald D Vale. 2017. “RNA Phase Transitions in Repeat Expansion Disorders.” Nature 546 (7657): 243–47. https://doi.org/10.1038/nature22386.

Jakobson, Christopher M, Edward Y Kim, Marilyn F Slininger, Alex Chien, and Danielle Tullman-Ercek. 2015. “Localization of Proteins to the 1,2-Propanediol Utilization Microcompartment by Non-Native Signal Sequences Is Mediated by a Common Hydrophobic Motif.” The Journal of Biological Chemistry 290 (40): 24519–33. https://doi.org/10.1074/jbc.M115.651919.

Joesaar, Alex, Shuo Yang, Bas Bögels, Ardjan van der Linden, Pascal Pieters, B. V.V.S.Pavan Kumar, Neil Dalchau, Andrew Phillips, Stephen Mann, and Tom F.A. de Greef. 2019. “DNA-Based Communication in Populations of Synthetic Protocells.” Nature Nanotechnology 14 (4): 369–78. https://doi.org/10.1038/s41565-019-0399-9.

Jones, Angharad R., Manuel Forero-Vargas, Simon P. Withers, Richard S. Smith, Jan Traas, Walter Dewitte, and James A.H. Murray. 2017. “Cell-Size Dependent Progression of the Cell Cycle Creates Homeostasis and Flexibility of Plant Cell Size.” Nature Communications 8 (April): 15060–15060. https://doi.org/10.1038/ncomms15060.

Jun, Suckjoon. 2015. “Chromosome, Cell Cycle, and Entropy.” Biophysical Journal. Biophysical Society. https://doi.org/10.1016/j.bpj.2014.12.032.

Karig, David, K. Michael Martini, Ting Lu, Nicholas A. DeLateur, Nigel Goldenfeld, and Ron Weiss. 2018. “Stochastic Turing Patterns in a Synthetic Bacterial Population.” Proceedings of the National Academy of Sciences of the United States of America 115 (26): 6572–77. https://doi.org/10.1073/pnas.1720770115.

Kerfeld, Cheryl A., and Onur Erbilgin. 2015. “Bacterial Microcompartments and the Modular Construction of Microbial Metabolism.” Trends in Microbiology. Elsevier Ltd.

43

https://doi.org/10.1016/j.tim.2014.10.003.

Kerfeld, Cheryl A., and Matthew R. Melnicki. 2016. “Assembly, Function and Evolution of Cyanobacterial Carboxysomes.” Current Opinion in Plant Biology. Elsevier Ltd. https://doi.org/10.1016/j.pbi.2016.03.009.

Kerfeld, Cheryl A. 2017. “A Bioarchitectonic Approach to the Modular Engineering of Metabolism.” Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences 372 (1730). https://doi.org/10.1098/rstb.2016.0387.

Kerfeld, Cheryl A, Clement Aussignargues, Jan Zarzycki, Fei Cai, and Markus Sutter. 2018. “Bacterial Microcompartments.” Nature Reviews. Microbiology 16 (5): 277–90. https://doi.org/10.1038/nrmicro.2018.10.

Kim, Edward Y, and Danielle Tullman-Ercek. 2014. “A Rapid Flow Cytometry Assay for the Relative Quantification of Protein Encapsulation into Bacterial Microcompartments.” Biotechnology Journal 9 (3): 348–54. https://doi.org/10.1002/biot.201300391.

Kim, Eugene, Jacob Kerssemakers, Indra A Shaltiel, Christian H Haering, and Cees Dekker. 2020. “DNA- Loop Extruding Condensin Complexes Can Traverse One Another.” Nature 579 (7799): 438–42. https://doi.org/10.1038/s41586-020-2067-5.

Kim, Wookhyun, and Elliot L. Chaikof. 2010. “Recombinant Elastin-Mimetic Biomaterials: Emerging Applications in Medicine.” Advanced Drug Delivery Reviews. Elsevier. https://doi.org/10.1016/j.addr.2010.04.007.

Klumpp, Jochen, and Thilo M. Fuchs. 2007. “Identification of Novel Genes in Genomic Islands That Contribute to Salmonella Typhimurium Replication in Macrophages.” Microbiology 153 (4): 1207–20. https://doi.org/10.1099/mic.0.2006/004747-0.

Koepnick, Brian, Jeff Flatten, Tamir Husain, Alex Ford, Daniel-Adriano Silva, Matthew J Bick, Aaron Bauer, et al. 2019. “De Novo Protein Design by Citizen Scientists.” Nature 570 (7761): 390–94. https://doi.org/10.1038/s41586-019-1274-4.

Kohyama, Shunshi, Natsuhiko Yoshinaga, Miho Yanagisawa, Kei Fujiwara, and Nobuhide Doi. 2019. “Cell- Sized Confinement Controls Generation and Stability of a Protein Wave for Spatiotemporal Regulation in Cells.” ELife 8 (July). https://doi.org/10.7554/eLife.44591.

44

Kretschmer, Simon, Katja Zieske, and Petra Schwille. 2017. “Large-Scale Modulation of Reconstituted Min Protein Patterns and Gradients by Defined Mutations in MinE’s Membrane Targeting Sequence.” PloS One 12 (6): e0179582. https://doi.org/10.1371/journal.pone.0179582.

Kumar, B. V.V.S.Pavan, Avinash J. Patil, and Stephen Mann. 2018. “Enzyme-Powered Motility in Buoyant Organoclay/DNA Protocells.” Nature Chemistry 10 (11): 1154–63. https://doi.org/10.1038/s41557-018- 0119-3.

Künzle, Matthias, Johanna Mangler, Marcel Lach, and Tobias Beck. 2018. “Peptide-Directed Encapsulation of Inorganic Nanoparticles into Protein Containers.” Nanoscale 10 (48): 22917–26. https://doi.org/10.1039/c8nr06236f.

Kurokawa, Chikako, Kei Fujiwara, Masamune Morita, Ibuki Kawamata, Yui Kawagishi, Atsushi Sakai, Yoshihiro Murayama, et al. 2017. “DNA Cytoskeleton for Stabilizing Artificial Cells.” Proceedings of the National Academy of Sciences of the United States of America 114 (28): 7228–33. https://doi.org/10.1073/pnas.1702208114.

Lagoutte, Priscillia, Charlotte Mignon, Gustavo Stadthagen, Supanee Potisopon, Stéphanie Donnat, Jan Mast, Adrien Lugari, and Bettina Werle. 2018. “Simultaneous Surface Display and Cargo Loading of Encapsulin Nanocompartments and Their Use for Rational Vaccine Design.” Vaccine 36 (25): 3622–28. https://doi.org/10.1016/j.vaccine.2018.05.034.

Larsen, Rachel A., Christina Cusumano, Akina Fujioka, Grace Lim-Fong, Paula Patterson, and Joe Pogliano. 2007. “Treadmilling of a Prokaryotic Tubulin-like Protein, TubZ, Required for Plasmid Stability in Bacillus Thuringiensis.” Genes and Development 21 (11): 1340–52. https://doi.org/10.1101/gad.1546107.

Lassila, Jonathan K, Susan L Bernstein, James N Kinney, Seth D Axen, and Cheryl A Kerfeld. 2014. “Assembly of Robust Bacterial Microcompartment Shells Using Building Blocks from an Organelle of Unknown Function.” Journal of Molecular Biology 426 (11): 2217–28. https://doi.org/10.1016/j.jmb.2014.02.025.

Last, Mart G.F., Siddharth Deshpande, and Cees Dekker. 2020. “PH-Controlled Coacervate-Membrane Interactions within Liposomes.” ACS Nano, April. https://doi.org/10.1021/acsnano.9b10167.

Lau, Yu Heng, Tobias W. Giessen, Wiggert J. Altenburg, and Pamela A. Silver. 2018. “Prokaryotic Nanocompartments Form Synthetic Organelles in a Eukaryote.” Nature Communications 9 (1): 1311.

45

https://doi.org/10.1038/s41467-018-03768-x.

Lawrence, Andrew D, Stefanie Frank, Sarah Newnham, Matthew J Lee, Ian R Brown, Wei-Feng Xue, Michelle L Rowe, et al. 2014. “Solution Structure of a Bacterial Microcompartment Targeting Peptide and Its Application in the Construction of an Ethanol Bioreactor.” ACS Synthetic Biology 3 (7): 454–65. https://doi.org/10.1021/sb4001118.

Lee, Keel Yong, Sung Jin Park, Keon Ah Lee, Se Hwan Kim, Heeyeon Kim, Yasmine Meroz, L. Mahadevan, et al. 2018. “Photosynthetic Artificial Organelles Sustain and Control ATP-Dependent Reactions in a Protocellular System.” Nature Biotechnology 36 (6): 530–35. https://doi.org/10.1038/nbt.4140.

Lee, Matthew J, David J Palmer, and Martin J Warren. 2019. “Biotechnological Advances in Bacterial Microcompartment Technology.” Trends in Biotechnology 37 (3): 325–36. https://doi.org/10.1016/j.tibtech.2018.08.006.

Lee, Tek Hyung, Timothy S. Carpenter, Patrik D’haeseleer, David F. Savage, and Mimi C. Yung. 2020. “Encapsulin Carrier Proteins for Enhanced Expression of Antimicrobial Peptides.” Biotechnology and Bioengineering 117 (3): 603–13. https://doi.org/10.1002/bit.27222.

Leger, Michelle M., Markéta Petru, Vojtěch Žárský, Laura Eme, Čestmír Vlček, Tommy Harding, B. Franz Lang, Marek Eliáš, Pavel Doležal, and Andrew J. Roger. 2015. “An Ancestral Bacterial Division System Is Widespread in Eukaryotic Mitochondria.” Proceedings of the National Academy of Sciences of the United States of America 112 (33): 10239–46. https://doi.org/10.1073/pnas.1421392112.

Lentini, Roberta, Noël Yeh Martín, Michele Forlin, Luca Belmonte, Jason Fontana, Michele Cornella, Laura Martini, et al. 2017. “Two-Way Chemical Communication between Artificial and Natural Cells.” ACS Central Science 3 (2): 117–23. https://doi.org/10.1021/acscentsci.6b00330.

Lentini, Roberta, Silvia Perez Santero, Fabio Chizzolini, Dario Cecchi, Jason Fontana, Marta Marchioretto, Cristina Del Bianco, et al. 2014. “Integrating Artificial with Natural Cells to Translate Chemical Messages That Direct E. Coli Behaviour.” Nature Communications 5 (May): 1–6. https://doi.org/10.1038/ncomms5012.

Li, Pilong, Sudeep Banjade, Hui Chun Cheng, Soyeon Kim, Baoyu Chen, Liang Guo, Marc Llaguno, et al. 2012. “Phase Transitions in the Assembly of Multivalent Signalling Proteins.” Nature 483 (7389): 336–40. https://doi.org/10.1038/nature10879.

Li, Yifeng. 2015. “Split-Inteins and Their Bioapplications.” Biotechnology Letters 37 (11): 2121–37. 46

https://doi.org/10.1007/s10529-015-1905-2.

Li, Zhengda, Shixuan Liu, and Qiong Yang. 2017. “Incoherent Inputs Enhance the Robustness of Biological Oscillators.” Cell Systems 5 (1): 72-81.e4. https://doi.org/10.1016/j.cels.2017.06.013.

Liang, Mingzhi, Stefanie Frank, Heinrich Lünsdorf, Martin J Warren, and Michael B Prentice. 2017. “Bacterial Microcompartment-Directed Polyphosphate Kinase Promotes Stable Polyphosphate Accumulation in E. Coli.” Biotechnology Journal 12 (3). https://doi.org/10.1002/biot.201600415.

Lim, Hoong Chuin, Ivan Vladimirovich Surovtsev, Bruno Gabriel Beltran, Fang Huang, Jörg Bewersdorf, and Christine Jacobs-Wagner. 2014. “Evidence for a DNA-Relay Mechanism in ParABS-Mediated Chromosome Segregation.” ELife 3 (3): e02758. https://doi.org/10.7554/eLife.02758.

Litschel, Thomas, Beatrice Ramm, Roel Maas, Michael Heymann, and Petra Schwille. 2018. “Beating Vesicles: Encapsulated Protein Oscillations Cause Dynamic Membrane Deformations.” Angewandte Chemie International Edition 57 (50): 16286–90. https://doi.org/10.1002/anie.201808750.

Long, M Scott, Clinton D Jones, Marcus R Helfrich, Lauren K Mangeney-Slavin, and Christine D Keating. 2005. “Dynamic Microcompartmentation in Synthetic Cells.” Proceedings of the National Academy of Sciences of the United States of America 102 (17): 5920–25. https://doi.org/10.1073/pnas.0409333102.

Loose, Martin, Elisabeth Fischer-Friedrich, Christoph Herold, Karsten Kruse, and Petra Schwille. 2011. “Min Protein Patterns Emerge from Rapid Rebinding and Membrane Interaction of MinE.” Nature Structural & Molecular Biology 18 (5): 577–83. https://doi.org/10.1038/nsmb.2037.

Loose, Martin, Elisabeth Fischer-Friedrich, Jonas Ries, Karsten Kruse, and Petra Schwille. 2008. “Spatial Regulators for Bacterial Cell Division Self-Organize into Surface Waves in Vitro.” Science (New York, N.Y.) 320 (5877): 789–92. https://doi.org/10.1126/science.1154413.

Love, Celina, Jan Steinkühler, David T Gonzales, Naresh Yandrapalli, Tom Robinson, Rumiana Dimova, and T-Y Dora Tang. 2020. “Reversible PH-Responsive Coacervate Formation in Lipid Vesicles Activates Dormant Enzymatic Reactions.” Angewandte Chemie (International Ed. in English) 59 (15): 5950–57. https://doi.org/10.1002/anie.201914893.

Lutkenhaus, Joe. 2007. “Assembly Dynamics of the Bacterial MinCDE System and Spatial Regulation of the Z Ring.” Annual Review of Biochemistry 76 (1): 539–62. https://doi.org/10.1146/annurev.biochem.75.103004.142652.

47

M, Ashburner, Golic KG, and Hawley RS. 2005. “Drosophila: A Laboratory Handbook., 2nd Ed.” Cold Spring Harbor Laboratory Press, 164.

Majumder, Sagardip, Jonathan Garamella, Ying Lin Wang, Maxwell Denies, Vincent Noireaux, and Allen P. Liu. 2017. “Cell-Sized Mechanosensitive and Biosensing Compartment Programmed with DNA.” Chemical Communications 53 (53): 7349–52. https://doi.org/10.1039/c7cc03455e.

Majumder, Sagardip, and Allen P Liu. 2017. “Bottom-up Synthetic Biology: Modular Design for Making Artificial Platelets.” Physical Biology 15 (1): 013001. https://doi.org/10.1088/1478-3975/aa9768.

Majumder, Sagardip, Patrick T. Willey, Maxwell S. DeNies, Allen P. Liu, and Gant Luxton. 2019. “A Synthetic Biology Platform for the Reconstitution and Mechanistic Dissection of LINC Complex Assembly.” Journal of Cell Science 132 (4). https://doi.org/10.1242/jcs.219451.

Majumder, Sagardip, Nadab Wubshet, and Allen P Liu. 2019. “Encapsulation of Complex Solutions Using Droplet Microfluidics towards the Synthesis of Artificial Cells.” Journal of Micromechanics and Microengineering 29 (8): 083001. https://doi.org/10.1088/1361-6439/AB2377.

Malay, Ali D, Naoyuki Miyazaki, Artur Biela, Soumyananda Chakraborti, Karolina Majsterkiewicz, Izabela Stupka, Craig S Kaplan, et al. 2019. “An Ultra-Stable Gold-Coordinated Protein Cage Displaying Reversible Assembly.” Nature 569 (7756): 438–42. https://doi.org/10.1038/s41586-019-1185-4.

Markow, T. A., S. Beall, and L. M. Matzkin. 2009. “Egg Size, Embryonic Development Time and Ovoviviparity in Drosophila Species.” Journal of Evolutionary Biology 22 (2): 430–34. https://doi.org/10.1111/j.1420-9101.2008.01649.x.

Martín, Laura, Emilio Castro, Artur Ribeiro, Matilde Alonso, and J. Carlos Rodríguez-Cabello. 2012. “Temperature-Triggered Self-Assembly of Elastin-like Block Co-Recombinamers: The Controlled Formation of Micelles and Vesicles in an Aqueous Medium.” Biomacromolecules 13 (2): 293–98. https://doi.org/10.1021/bm201436y.

Martin, William. 2010. “Evolutionary Origins of Metabolic Compartmentalization in Eukaryotes.” Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences 365 (1541): 847– 55. https://doi.org/10.1098/rstb.2009.0252.

Martos, Ariadna, Zdenek Petrasek, and Petra Schwille. 2013. “Propagation of MinCDE Waves on Free- Standing Membranes.” Environmental Microbiology 15 (12): 3319–26. https://doi.org/10.1111/1462- 2920.12295. 48

Matamoro-Vidal, Alexis, Isaac Salazar-Ciudad, and David Houle. 2015. “Making Quantitative Morphological Variation from Basic Developmental Processes: Where Are We? The Case of the Drosophila Wing.” Developmental Dynamics 244 (9): 1058–73. https://doi.org/10.1002/dvdy.24255.

Maurizi, M R. 1992. “Proteases and Protein Degradation in Escherichia Coli.” Experientia 48 (2): 178– 201. https://doi.org/10.1007/bf01923511.

Mayer, Matthias J, Rokas Juodeikis, Ian R Brown, Stefanie Frank, David J Palmer, Evelyne Deery, David M Beal, Wei-Feng Xue, and Martin J Warren. 2016. “Effect of Bio-Engineering on Size, Shape, Composition and Rigidity of Bacterial Microcompartments.” Scientific Reports 6 (November): 36899. https://doi.org/10.1038/srep36899.

McHugh, Colleen A, Juan Fontana, Daniel Nemecek, Naiqian Cheng, Anastasia A Aksyuk, J Bernard Heymann, Dennis C Winkler, et al. 2014. “A Virus Capsid-like Nanocompartment That Stores Iron and Protects Bacteria from Oxidative Stress.” The EMBO Journal 33 (17): 1896–1911. https://doi.org/10.15252/embj.201488566.

Mizuuchi, Kiyoshi, and Anthony G Vecchiarelli. 2018. “Mechanistic Insights of the Min Oscillator via Cell- Free Reconstitution and Imaging.” Physical Biology 15 (3): 031001. https://doi.org/10.1088/1478- 3975/aa9e5e.

Møller-Jensen, Jakob, Jonas Borch, Mette Dam, Rasmus B. Jensen, Peter Roepstorff, and Kenn Gerdes. 2003. “Bacterial Mitosis: ParM of Plasmid R1 Moves Plasmid DNA by an Actin-like Insertional Polymerization Mechanism.” Molecular Cell 12 (6): 1477–87. https://doi.org/10.1016/S1097- 2765(03)00451-9.

Molliex, Amandine, Jamshid Temirov, Jihun Lee, Maura Coughlin, Anderson P Kanagaraj, Hong Joo Kim, Tanja Mittag, and J Paul Taylor. 2015. “Phase Separation by Low Complexity Domains Promotes Stress Granule Assembly and Drives Pathological Fibrillization.” Cell 163 (1): 123–33. https://doi.org/10.1016/j.cell.2015.09.015.

Monahan, Leigh G, Andrew T F Liew, Amy L Bottomley, and Elizabeth J Harry. 2014. “Division Site Positioning in Bacteria: One Size Does Not Fit All.” Frontiers in Microbiology 5 (FEB): 19. https://doi.org/10.3389/fmicb.2014.00019.

Monterroso, Begoña, Silvia Zorrilla, Marta Sobrinos-Sanguino, Miguel A Robles-Ramos, Marina López- Álvarez, William Margolin, Christine D Keating, and Germán Rivas. 2019. “Bacterial FtsZ Protein Forms

49

Phase-separated Condensates with Its Nucleoid-associated Inhibitor SlmA.” EMBO Reports 20 (1). https://doi.org/10.15252/embr.201845946.

Moon, Hyojin, Jisu Lee, Junseon Min, and Sebyung Kang. 2014. “Developing Genetically Engineered Encapsulin Protein Cage Nanoparticles as a Targeted Delivery Nanoplatform.” Biomacromolecules 15 (10): 3794–3801. https://doi.org/10.1021/bm501066m.

Murphy, L D, and S B Zimmerman. 1997. “Stabilization of Compact Spermidine Nucleoids from Escherichia Coli under Crowded Conditions: Implications for in Vivo Nucleoid Structure.” Journal of Structural Biology 119 (3): 336–46. https://doi.org/10.1006/jsbi.1997.3884.

Nakajima, Masato, Keiko Imai, Hiroshi Ito, Taeko Nishiwaki, Yoriko Murayama, Hideo Iwasaki, Tokitaka Oyama, and Takao Kondo. 2005. “Reconstitution of Circadian Oscillation of Cyanobacterial KaiC Phosphorylation in Vitro.” Science 308 (5720): 414–15. https://doi.org/10.1126/science.1108451.

Nelson, D E, and K D Young. 2000. “Penicillin Binding Protein 5 Affects Cell Diameter, Contour, and Morphology of Escherichia Coli.” Journal of Bacteriology 182 (6): 1714–21. https://doi.org/10.1128/jb.182.6.1714-1721.2000.

Nichols, Robert J., Caleb Cassidy-Amstutz, Thawatchai Chaijarasphong, and David F. Savage. 2017. “Encapsulins: Molecular Biology of the Shell.” Critical Reviews in Biochemistry and Molecular Biology. Taylor and Francis Ltd. https://doi.org/10.1080/10409238.2017.1337709.

Noireaux, Vincent, and Allen P. Liu. 2020. “The New Age of Cell-Free Biology.” Annual Review of Biomedical Engineering 22 (1). https://doi.org/10.1146/annurev-bioeng-092019-111110.

Oberholzer, Thomas, Roger Wick, Pier Luigi Luisi, and Christof K. Biebricher. 1995. “Enzymatic RNA Replication in Self-Reproducing Vesicles: An Approach to a Minimal Cell.” Biochemical and Biophysical Research Communications. https://doi.org/10.1006/bbrc.1995.1180.

Odijk, Theo. 1998. “Osmotic Compaction of Supercoiled DNA into a Bacterial Nucleoid.” Biophysical Chemistry 73 (1–2): 23–29. https://doi.org/10.1016/S0301-4622(98)00115-X.

Otrin, Lado, Nika Marušič, Claudia Bednarz, Tanja Vidaković-Koch, Ingo Lieberwirth, Katharina Landfester, and Kai Sundmacher. 2017. “Toward Artificial Mitochondrion: Mimicking Oxidative Phosphorylation in Polymer and Hybrid Membranes.” Nano Letters 17 (11): 6816–21. https://doi.org/10.1021/acs.nanolett.7b03093.

50

Parsons, Joshua B, Stefanie Frank, David Bhella, Mingzhi Liang, Michael B Prentice, Daniel P Mulvihill, and Martin J Warren. 2010. “Synthesis of Empty Bacterial Microcompartments, Directed Organelle Protein Incorporation, and Evidence of Filament-Associated Organelle Movement.” Molecular Cell 38 (2): 305–15. https://doi.org/10.1016/j.molcel.2010.04.008.

Pautot, Sophie, Barbara J. Frisken, and D. A. Weitz. 2003. “Production of Unilamellar Vesicles Using an Inverted Emulsion.” Langmuir 19 (7): 2870–79. https://doi.org/10.1021/la026100v.

Pazos, Manuel, Paolo Natale, and Miguel Vicente. 2013. “A Specific Role for the ZipA Protein in Cell Division: Stabilization of the FtsZ Protein.” The Journal of Biological Chemistry 288 (5): 3219–26. https://doi.org/10.1074/jbc.M112.434944.

Peters, Ruud J. R. W., Maïté Marguet, Sébastien Marais, Marco W. Fraaije, Jan C. M. van Hest, and Sébastien Lecommandoux. 2014. “Cascade Reactions in Multicompartmentalized Polymersomes.” Angewandte Chemie International Edition 53 (1): 146–50. https://doi.org/10.1002/anie.201308141.

Petit, Elsa, W. Greg LaTouf, Maddalena V. Coppi, Thomas A. Warnick, Devin Currie, Igor Romashko, Supriya Deshpande, et al. 2013. “Involvement of a Bacterial Microcompartment in the Metabolism of Fucose and Rhamnose by Clostridium Phytofermentans.” PLoS ONE 8 (1). https://doi.org/10.1371/journal.pone.0054337.

Peyret, Ariane, Emmanuel Ibarboure, Natassa Pippa, and Sebastien Lecommandoux. 2017. “Liposomes in Polymersomes: Multicompartment System with Temperature-Triggered Release.” Langmuir 33 (28): 7079–85. https://doi.org/10.1021/acs.langmuir.7b00655.

Pitts, Alison C., Laura R. Tuck, Alexandra Faulds-Pain, Richard J. Lewis, and Jon Marles-Wright. 2012. “Structural Insight into the Clostridium Difficile Ethanolamine Utilisation Microcompartment.” PLoS ONE 7 (10). https://doi.org/10.1371/journal.pone.0048360.

Planamente, Sara, and Stefanie Frank. 2019. “Bio-Engineering of Bacterial Microcompartments: A Mini Review.” Biochemical Society Transactions 47 (3): 765–77. https://doi.org/10.1042/BST20170564.

Plegaria, Jefferson S, and Cheryl A Kerfeld. 2018. “Engineering Nanoreactors Using Bacterial Microcompartment Architectures.” Current Opinion in Biotechnology 51 (June): 1–7. https://doi.org/10.1016/j.copbio.2017.09.005.

Putri, Rindia M, Carolina Allende-Ballestero, Daniel Luque, Robin Klem, Katerina-Asteria Rousou, Aijie Liu, Christoph H-H Traulsen, et al. 2017. “Structural Characterization of Native and Modified Encapsulins 51

as Nanoplatforms for in Vitro Catalysis and Cellular Uptake.” ACS Nano 11 (12): 12796–804. https://doi.org/10.1021/acsnano.7b07669.

Rahmanpour, Rahman, and Timothy D H Bugg. 2013. “Assembly in Vitro of Rhodococcus Jostii RHA1 Encapsulin and Peroxidase DypB to Form a Nanocompartment.” The FEBS Journal 280 (9): 2097–2104. https://doi.org/10.1111/febs.12234.

Ramm, Beatrice, Philipp Glock, and Petra Schwille. 2018. “In Vitro Reconstitution of Self-Organizing Protein Patterns on Supported Lipid Bilayers.” Journal of Visualized Experiments 2018 (137). https://doi.org/10.3791/58139.

Raskin, David M., and Piet A.J. De Boer. 1999. “Rapid Pole-to-Pole Oscillation of a Protein Required for Directing Division to the Middle of Escherichia Coli.” Proceedings of the National Academy of Sciences of the United States of America 96 (9): 4971–76. https://doi.org/10.1073/pnas.96.9.4971.

Rico, Ana Isabel, Marcin Krupka, and Miguel Vicente. 2013. “In the Beginning, Escherichia Coli Assembled the Proto-Ring: An Initial Phase of Division.” The Journal of Biological Chemistry 288 (29): 20830–36. https://doi.org/10.1074/jbc.R113.479519.

Saha, Shambaditya, Christoph A. Weber, Marco Nousch, Omar Adame-Arana, Carsten Hoege, Marco Y. Hein, Erin Osborne-Nishimura, et al. 2016. “Polar Positioning of Phase-Separated Liquid Compartments in Cells Regulated by an MRNA Competition Mechanism.” Cell 166 (6): 1572-1584.e16. https://doi.org/10.1016/j.cell.2016.08.006.

Schaffter, Samuel W., and Rebecca Schulman. 2019. “Building in Vitro Transcriptional Regulatory Networks by Successively Integrating Multiple Functional Circuit Modules.” Nature Chemistry 11 (9): 829–38. https://doi.org/10.1038/s41557-019-0292-z.

Schramm, Antoine, Philippe Lieutaud, Stefano Gianni, Sonia Longhi, and Christophe Bignon. 2017. “InSiDDe: A Server for Designing Artificial Disordered Proteins.” International Journal of Molecular Sciences 19 (1). https://doi.org/10.3390/ijms19010091.

Schulz, Matthias, and Wolfgang H. Binder. 2015. “Mixed Hybrid Lipid/Polymer Vesicles as a Novel Membrane Platform.” Macromolecular Rapid Communications 36 (23): 2031–41. https://doi.org/10.1002/marc.201500344.

Schuster, Benjamin S, Ellen H Reed, Ranganath Parthasarathy, Craig N Jahnke, Reese M Caldwell, Jessica G Bermudez, Holly Ramage, Matthew C Good, and Daniel A Hammer. 2018. “Controllable Protein Phase 52

Separation and Modular Recruitment to Form Responsive Membraneless Organelles.” Nature Communications 9 (1): 2985. https://doi.org/10.1038/s41467-018-05403-1.

Schweizer, Jakob, Martin Loose, Mike Bonny, Karsten Kruse, Ingolf Mönch, and Petra Schwille. 2012. “Geometry Sensing by Self-Organized Protein Patterns.” Proceedings of the National Academy of Sciences of the United States of America 109 (38): 15283–88. https://doi.org/10.1073/pnas.1206953109.

Schwille, Petra, Joachim Spatz, Katharina Landfester, Eberhard Bodenschatz, Stephan Herminghaus, Victor Sourjik, Tobias J. Erb, et al. 2018. “MaxSynBio: Avenues Towards Creating Cells from the Bottom Up.” Angewandte Chemie - International Edition. Wiley-VCH Verlag. https://doi.org/10.1002/anie.201802288.

Sezonov, Guennadi, Danièle Joseleau-Petit, and Richard D’Ari. 2007. “Escherichia Coli Physiology in Luria-Bertani Broth.” Journal of Bacteriology 189 (23): 8746–49. https://doi.org/10.1128/JB.01368-07.

Shimizu, Yoshihiro, Akio Inoue, Yukihide Tomari, Tsutomu Suzuki, Takashi Yokogawa, Kazuya Nishikawa, and Takuya Ueda. 2001. “Cell-Free Translation Reconstituted with Purified Components.” Nature Biotechnology 19 (8): 751–55. https://doi.org/10.1038/90802.

Shin, Jonghyeon, Paul Jardine, and Vincent Noireaux. 2012. “Genome Replication, Synthesis, and Assembly of the Bacteriophage T7 in a Single Cell-Free Reaction.” ACS Synthetic Biology 1 (9): 408–13. https://doi.org/10.1021/sb300049p.

Shinoda, Takehiro, Naoko Shinya, Kaori Ito, Yoshiko Ishizuka-Katsura, Noboru Ohsawa, Takaho Terada, Kunio Hirata, et al. 2016. “Cell-Free Methods to Produce Structurally Intact Mammalian Membrane Proteins.” Scientific Reports 6 (1): 1–15. https://doi.org/10.1038/srep30442.

Sigmund, Felix, Christoph Massner, Philipp Erdmann, Anja Stelzl, Hannes Rolbieski, Mitul Desai, Sarah Bricault, et al. 2018. “Bacterial Encapsulins as Orthogonal Compartments for Mammalian Cell Engineering.” Nature Communications 9 (1): 1990. https://doi.org/10.1038/s41467-018-04227-3.

Sigmund, Felix, Susanne Pettinger, Massimo Kube, Fabian Schneider, Martina Schifferer, Steffen Schneider, Maria V Efremova, et al. 2019. “Iron-Sequestering Nanocompartments as Multiplexed Electron Microscopy Gene Reporters.” ACS Nano 13 (7): 8114–23. https://doi.org/10.1021/acsnano.9b03140.

Slininger Lee, Marilyn F, Christopher M Jakobson, and Danielle Tullman-Ercek. 2017. “Evidence for Improved Encapsulated Pathway Behavior in a Bacterial Microcompartment through Shell Protein 53

Engineering.” ACS Synthetic Biology 6 (10): 1880–91. https://doi.org/10.1021/acssynbio.7b00042.

Sonotaki, Seiichi, Taku Takami, Keiichi Noguchi, Masafumi Odaka, Masafumi Yohda, and Yoshihiko Murakami. 2017. “Successful PEGylation of Hollow Encapsulin Nanoparticles from Rhodococcus Erythropolis N771 without Affecting Their Disassembly and Reassembly Properties.” Biomaterials Science 5 (6): 1082–89. https://doi.org/10.1039/c7bm00207f.

Staller, Max V., Alex S. Holehouse, Devjanee Swain-Lenz, Rahul K. Das, Rohit V. Pappu, and Barak A. Cohen. 2018. “A High-Throughput Mutational Scan of an Intrinsically Disordered Acidic Transcriptional Activation Domain.” Cell Systems 6 (4): 444-455.e6. https://doi.org/10.1016/j.cels.2018.01.015.

Sun, Meng, Zhengda Li, Shiyuan Wang, Gembu Maryu, and Qiong Yang. 2019. “Building Dynamic Cellular Machineries in Droplet-Based Artificial Cells with Single-Droplet Tracking and Analysis.” Analytical Chemistry 91 (15): 9813–18. https://doi.org/10.1021/acs.analchem.9b01481.

Surovtsev, Ivan V., and Christine Jacobs-Wagner. 2018. “Subcellular Organization: A Critical Feature of Bacterial Cell Replication.” Cell. Cell Press. https://doi.org/10.1016/j.cell.2018.01.014.

Sutter, Markus, Daniel Boehringer, Sascha Gutmann, Susanne Günther, David Prangishvili, Martin J. Loessner, Karl O. Stetter, Eilika Weber-Ban, and Nenad Ban. 2008. “Structural Basis of Enzyme Encapsulation into a Bacterial Nanocompartment.” Nature Structural and Molecular Biology 15 (9): 939– 47. https://doi.org/10.1038/nsmb.1473.

Sutter, Markus, Basil Greber, Clement Aussignargues, and Cheryl A Kerfeld. 2017. “Assembly Principles and Structure of a 6.5-MDa Bacterial Microcompartment Shell.” Science (New York, N.Y.) 356 (6344): 1293–97. https://doi.org/10.1126/science.aan3289.

Swank, Zoe, Nadanai Laohakunakorn, and Sebastian J. Maerkl. 2019. “Cell-Free Gene-Regulatory Network Engineering with Synthetic Transcription Factors.” Proceedings of the National Academy of Sciences of the United States of America 116 (13): 5892–5901. https://doi.org/10.1073/pnas.1816591116.

Szwedziak, Piotr, and Debnath Ghosal. 2017. “FtsZ-Ring Architecture and Its Control by MinCD.” Sub- Cellular Biochemistry 84: 213–44. https://doi.org/10.1007/978-3-319-53047-5_7.

Szwedziak, Piotr, Qing Wang, Tanmay A M Bharat, Matthew Tsim, and Jan Löwe. 2014. “Architecture of the Ring Formed by the Tubulin Homologue FtsZ in Bacterial Cell Division.” ELife 3 (December): e04601. https://doi.org/10.7554/eLife.04601. 54

Tamura, Akio, Yosuke Fukutani, Taku Takami, Motoko Fujii, Yuki Nakaguchi, Yoshihiko Murakami, Keiichi Noguchi, Masafumi Yohda, and Masafumi Odaka. 2015. “Packaging Guest Proteins into the Encapsulin Nanocompartment from Rhodococcus Erythropolis N771.” Biotechnology and Bioengineering 112 (1): 13–20. https://doi.org/10.1002/bit.25322.

Tanaka, Shiho, Cheryl A. Kerfeld, Michael R. Sawaya, Fei Cai, Sabine Heinhorst, Gordon C. Cannon, and Todd O. Yeates. 2008. “Atomic-Level Models of the Bacterial Carboxysome Shell.” Science 319 (5866): 1083–86. https://doi.org/10.1126/science.1151458.

Tayar, Alexandra M., Eyal Karzbrun, Vincent Noireaux, and Roy H. Bar-Ziv. 2017. “Synchrony and Pattern Formation of Coupled Genetic Oscillators on a Chip of Artificial Cells.” Proceedings of the National Academy of Sciences of the United States of America 114 (44): 11609–14. https://doi.org/10.1073/pnas.1710620114.

Thiennimitr, Parameth, Sebastian E Winter, Maria G Winter, Mariana N Xavier, Vladimir Tolstikov, Douglas L Huseby, Torsten Sterzenbach, Renée M Tsolis, John R Roth, and Andreas J Bäumler. 2011. “Intestinal Inflammation Allows Salmonella to Use Ethanolamine to Compete with the Microbiota.” Proceedings of the National Academy of Sciences of the United States of America 108 (42): 17480–85. https://doi.org/10.1073/pnas.1107857108.

Thompson, John F., and Arthur Landy. 1988. “Empirical Estimation of Protein-Induced DNA Bending Angles: Applications to λ Site-Specific Recombination Complexes.” Nucleic Acids Research 16 (20): 9687– 9705. https://doi.org/10.1093/nar/16.20.9687.

Ting, Claire S, Katharine H Dusenbury, Reid A Pryzant, Kathleen W Higgins, Catherine J Pang, Christie E Black, and Ellen M Beauchamp. 2015. “The Prochlorococcus Carbon Dioxide-Concentrating Mechanism: Evidence of Carboxysome-Associated Heterogeneity.” Photosynthesis Research 123 (1): 45–60. https://doi.org/10.1007/s11120-014-0038-0.

Toda, Satoshi, Lucas R. Blauch, Sindy K.Y. Tang, Leonardo Morsut, and Wendell A. Lim. 2018. “Programming Self-Organizing Multicellular Structures with Synthetic Cell-Cell Signaling.” Science 361 (6398): 156–62. https://doi.org/10.1126/science.aat0271.

Tracey, John C, Maricela Coronado, Tobias W Giessen, Maggie C Y Lau, Pamela A Silver, and Bess B Ward. 2019. “The Discovery of Twenty-Eight New Encapsulin Sequences, Including Three in Anammox Bacteria.” Scientific Reports 9 (1): 20122. https://doi.org/10.1038/s41598-019-56533-5.

55

Trantidou, T., M. S. Friddin, A. Salehi-Reyhani, O. Ces, and Y. Elani. 2018. “Droplet Microfluidics for the Construction of Compartmentalised Model Membranes.” Lab on a Chip. Royal Society of Chemistry. https://doi.org/10.1039/c8lc00028j.

Tsai, Tony Yu Chen, Sup Choi Yoon, Wenzhe Ma, Joseph R. Pomerening, Chao Tang, and James E. Ferrell. 2008. “Robust, Tunable Biological Oscillations from Interlinked Positive and Negative Feedback Loops.” Science 321 (5885): 126–39. https://doi.org/10.1126/science.1156951.

Turmo, Aiko, C Raul Gonzalez-Esquer, and Cheryl A Kerfeld. 2017. “Carboxysomes: Metabolic Modules for CO2 Fixation.” FEMS Microbiology Letters 364 (18). https://doi.org/10.1093/femsle/fnx176.

Uddin, Ismail, Stefanie Frank, Martin J Warren, and Richard W Pickersgill. 2018. “A Generic Self- Assembly Process in Microcompartments and Synthetic Protein Nanotubes.” Small (Weinheim an Der Bergstrasse, Germany) 14 (19): e1704020. https://doi.org/10.1002/smll.201704020.

Vargo, Kevin B., Ranganath Parthasarathy, and Daniel A. Hammer. 2012. “Self-Assembly of Tunable Protein Suprastructures from Recombinant Oleosin.” Proceedings of the National Academy of Sciences of the United States of America 109 (29): 11657–62. https://doi.org/10.1073/pnas.1205426109.

Vecchiarelli, Anthony G, Ling Chin Hwang, and Kiyoshi Mizuuchi. 2013. “Cell-Free Study of F Plasmid Partition Provides Evidence for Cargo Transport by a Diffusion-Ratchet Mechanism.” Proceedings of the National Academy of Sciences of the United States of America 110 (15): E1390-7. https://doi.org/10.1073/pnas.1302745110.

Vecchiarelli, Anthony G, Min Li, Michiyo Mizuuchi, Ling Chin Hwang, Yeonee Seol, Keir C Neuman, and Kiyoshi Mizuuchi. 2016. “Membrane-Bound MinDE Complex Acts as a Toggle Switch That Drives Min Oscillation Coupled to Cytoplasmic Depletion of MinD.” Proceedings of the National Academy of Sciences of the United States of America 113 (11): E1479-88. https://doi.org/10.1073/pnas.1600644113.

Vecchiarelli, Anthony G, Min Li, Michiyo Mizuuchi, and Kiyoshi Mizuuchi. 2014. “Differential Affinities of MinD and MinE to Anionic Phospholipid Influence Min Patterning Dynamics in Vitro.” Molecular Microbiology 93 (3): 453–63. https://doi.org/10.1111/mmi.12669.

Vecchiarelli, Anthony G, Keir C Neuman, and Kiyoshi Mizuuchi. 2014. “A Propagating ATPase Gradient Drives Transport of Surface-Confined Cellular Cargo.” Proceedings of the National Academy of Sciences of the United States of America 111 (13): 4880–85. https://doi.org/10.1073/pnas.1401025111.

Vogele, Kilian, Thomas Frank, Lukas Gasser, Marisa A. Goetzfried, Mathias W. Hackl, Stephan A. Sieber, 56

Friedrich C. Simmel, and Tobias Pirzer. 2018. “Towards Synthetic Cells Using Peptide-Based Reaction Compartments.” Nature Communications 9 (1): 1–7. https://doi.org/10.1038/s41467-018-06379-8.

Walpole, Joseph, Jason A. Papin, and Shayn M. Peirce. 2013. “Multiscale Computational Models of Complex Biological Systems.” Annual Review of Biomedical Engineering 15 (1): 137–54. https://doi.org/10.1146/annurev-bioeng-071811-150104.

Walter, J-C, J Dorignac, V Lorman, J Rech, J-Y Bouet, M Nollmann, J Palmeri, A Parmeggiani, and F Geniet. 2017. “Surfing on Protein Waves: Proteophoresis as a Mechanism for Bacterial Genome Partitioning.” Physical Review Letters 119 (2): 028101. https://doi.org/10.1103/PhysRevLett.119.028101.

Wang, Shangying, Kai Fan, Nan Luo, Yangxiaolu Cao, Feilun Wu, Carolyn Zhang, Katherine A. Heller, and Lingchong You. 2019. “Massive Computational Acceleration by Using Neural Networks to Emulate Mechanism-Based Biological Models.” Nature Communications 10 (1): 1–9. https://doi.org/10.1038/s41467-019-12342-y.

Wang, Xuejing, Hang Du, Zhao Wang, Wei Mu, and Xiaojun Han. 2020. “Versatile Phospholipid Assemblies for Functional Synthetic Cells and Artificial Tissues.” Advanced Materials. https://doi.org/10.1002/adma.202002635.

Weiss, Marian, Johannes Patrick Frohnmayer, Lucia Theresa Benk, Barbara Haller, Jan Willi Janiesch, Thomas Heitkamp, Michael Börsch, et al. 2018. “Sequential Bottom-up Assembly of Mechanically Stabilized Synthetic Cells by Microfluidics.” Nature Materials 17 (1): 89–95. https://doi.org/10.1038/NMAT5005.

Weitz, J. 2016. “Quantitative Viral Ecology: Dynamics of Viruses and Their Microbial Hosts.” Princeton University Press 55.

Weitz, Maximilian, Jongmin Kim, Korbinian Kapsner, Erik Winfree, Elisa Franco, and Friedrich C. Simmel. 2014. “Diversity in the Dynamical Behaviour of a Compartmentalized Programmable Biochemical Oscillator.” Nature Chemistry 6 (4): 295–302. https://doi.org/10.1038/nchem.1869.

Williams, Elsie M, Se Min Jung, Jennifer L Coffman, and Stefan Lutz. 2018. “Pore Engineering for Enhanced Mass Transport in Encapsulin Nanocompartments.” ACS Synthetic Biology 7 (11): 2514–17. https://doi.org/10.1021/acssynbio.8b00295.

Xu, Can, Shuo Hu, and Xiaoyuan Chen. 2016. “Artificial Cells: From Basic Science to Applications.” Biochemical Pharmacology 19 (9): 516–32. https://doi.org/10.1016/j.mattod.2016.02.020. 57

Yung, Mimi C, Feliza A Bourguet, Timothy S Carpenter, and Matthew A Coleman. 2017. “Re-Directing Bacterial Microcompartment Systems to Enhance Recombinant Expression of Lysis Protein E from Bacteriophage ΦX174 in Escherichia Coli.” Microbial Cell Factories 16 (1): 71. https://doi.org/10.1186/s12934-017-0685-x.

Zakeri, Bijan, Jacob O. Fierer, Emrah Celik, Emily C. Chittock, Ulrich Schwarz-Linek, Vincent T. Moy, and Mark Howarth. 2012. “Peptide Tag Forming a Rapid Covalent Bond to a Protein, through Engineering a Bacterial Adhesin.” Proceedings of the National Academy of Sciences of the United States of America 109 (12). https://doi.org/10.1073/pnas.1115485109.

Zhao, L., C. D. Kroenke, J. Song, D. Piwnica-Worms, J. J.H. Ackerman, and J. J. Neil. 2008. “Intracellular Water-Specific MR of Microbead-Adherent Cells: The HeLa Cell Intracellular Water Exchange Lifetime.” NMR in Biomedicine 21 (2): 159–64. https://doi.org/10.1002/nbm.1173.

Zhou, Huaijin, and Joe Lutkenhaus. 2003. “Membrane Binding by MinD Involves Insertion of Hydrophobic Residues within the C-Terminal Amphipathic Helix into the Bilayer.” Journal of Bacteriology 185 (15): 4326–35. https://doi.org/10.1128/JB.185.15.4326-4335.2003.

Zieske, Katja, and Petra Schwille. 2014. “Reconstitution of Self-Organizing Protein Gradients as Spatial Cues in Cell-Free Systems.” ELife 3 (October). https://doi.org/10.7554/eLife.03949.

Godino, Elisa, Jonás Noguera López, Ilias Zarguit, Anne Doerr, Mercedes Jimenez, Germán Rivas, and Christophe Danelon. 2020. “Cell-Free Biogenesis of Bacterial Division Proto-Rings That Can Constrict Liposomes.” Communications Biology 3 (1): 539. https://doi.org/10.1038/s42003-020-01258-9.

Göpfrich, Kerstin, Ilia Platzman, and Joachim P. Spatz. 2018. “Mastering Complexity: Towards Bottom-up Construction of Multifunctional Eukaryotic Synthetic Cells.” Trends in Biotechnology. Elsevier Ltd. https://doi.org/10.1016/j.tibtech.2018.03.008.

Schwille, Petra, Joachim Spatz, Katharina Landfester, Eberhard Bodenschatz, Stephan Herminghaus, Victor Sourjik, Tobias J. Erb, et al. 2018. “MaxSynBio: Avenues Towards Creating Cells from the Bottom Up.” Angewandte Chemie - International Edition. Wiley-VCH Verlag. https://doi.org/10.1002/anie.201802288.

Wang, Xuejing, Hang Du, Zhao Wang, Wei Mu, and Xiaojun Han. 2020. “Versatile Phospholipid Assemblies for Functional Synthetic Cells and Artificial Tissues.” Advanced Materials. https://doi.org/10.1002/adma.202002635.

58