Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Oct 1.
Published in final edited form as: Curr Opin Biotechnol. 2010 Jul 16;21(5):697–703. doi: 10.1016/j.copbio.2010.06.008

Update on designing and building minimal cells

Michael C Jewett 1, Anthony C Forster 2
PMCID: PMC2952674  NIHMSID: NIHMS217783  PMID: 20638265

Summary

Minimal cells comprise only the genes and biomolecular machinery necessary for basic life. Synthesizing minimal and minimized cells will improve understanding of core biology, enhance development of biotechnology strains of bacteria, and enable evolutionary optimization of natural and unnatural biopolymers. Design and construction of minimal cells is proceeding in two different directions: “top-down” reduction of bacterial genomes in vivo and “bottom-up” integration of DNA/RNA/protein/membrane syntheses in vitro. Major progress in the last 5 years has occurred in synthetic genomics, minimization of the Escherichia coli genome, sequencing of minimal bacterial endosymbionts, identification of essential genes, and integration of biochemical systems.

Introduction

Design-based engineering of biological systems (also known as synthetic biology) tests understanding of the living world and harnesses its diverse repertoire to solve society’s problems [1,2]. Ideally, an engineered system should be functionally robust and predictable. Yet these features are difficult to achieve when engineering biology [3] because of the poorly understood complexity of even the simplest single-celled organisms. An enticing way to simplify cellular complexity, test understanding and potentially facilitate engineering is to synthesize minimal cells [47]. Forster and Church reviewed plans of others to minimize small bacterial cells (in vivo “top-down” approach) [5] and proposed detailed plans for synthesizing a minimal cell from biomolecular parts (in vitro “bottom-up” approach) [4]. Here, we highlight progress, challenges and prospects since these two reviews.

New tools

Minimal cells require minimal genomes, and minimal genomes require design, construction and manipulation tools at an unprecedented scale. Great progress has been made in genome construction by the J. Craig Venter Institute (JCVI; Rockville, MD, USA). JCVI constructed the 582 kilobase pair (kbp) genome of Mycoplasma genitalium, the smallest known genome of a bacterium capable of independent growth [8]. This was done by commercial gene synthesis from oligodeoxyribonucleotides (oligos) and then stepwise assembly. Assemblies of up to quarter genomes were cloned in vitro in E. coli bacterial artificial chromosomes, while the final assembly used recombination in the yeast Saccharomyces cerevisiae [9]. JCVI further improved the technology by enzymatic assembly of genes in vitro [10] and by discovering that yeast has the remarkable capability of simultaneously recombining 25 overlapping DNA fragments to make the complete M. genitalium genome [11]. More recently, JCVI has developed methods for manipulating and cloning whole genomes in yeast [12] and shown synthesis of a larger 1.08 million base-pair M. mycoides JCVI-syn1.0 genome [13]. Though sequencing has become inexpensive, the costs of chemically synthesizing genes have leveled out at ~$0.50/bp which is prohibitive at the genome scale for typical researchers. More affordable genetic segments may be obtained from native genomes by restriction digestion or PCR-amplification, which may limit sequence design, or by improved methods for assembling genes from oligos [14,15].

In contrast to genome construction, non-viral genome design and manipulation are still primitive and certainly cannot be done from scratch. For example, substantial changes in whole bacterial genomes essentially have been limited to conservative deletions (see below and [16]), programming microbes for expression of the anti-malarial drug artemisinin has taken 150-person years of work [17], and coordinated over-expression of multiple proteins in a single cell is difficult to achieve [18]. Optimization and discovery of new designs will be helped by directed evolution technologies such as multiplex automated genome engineering (MAGE; [19]). MAGE generates genomic diversity in E. coli by parallel, oligo-directed, genomic modifications.

Top-down approach: in vivo reduction

Even the most highly reduced genome of M. genitalium contains 100 individually dispensable genes out of 528 annotated genes [20], so streamlining down to only essential genes is one route to minimal cells. So far, significant minimization has been carried out only in organisms with larger genomes such as E. coli (4,640 kbp; 4,434 genes) and Bacillus subtilis (4,216 kbp; 4,245 genes) aided by known sequences of closely-related genomes. Genome reduction by up to 30% has proven surprisingly successful for viability, genome stabilization [21], promoting growth [22] and enhancing recombinant protein production [23]. Rather than targeted deletion, genome reductions of up to 200 kbp can also result from experimental evolution [24].

The smallest minimized genomes from the top-down approach will likely be produced by minimizing the already smallest genome, that of M. genitalium. JCVI has been pursuing this plan in 6 ambitious steps:

  1. sequencing the M. genitalium genome [8] (and related genomes),

  2. defining its genes that are individually dispensable [20],

  3. re-synthesizing the genome from oligos [9,11],

  4. transplanting the synthetic (donor) genome into related Mycoplasma recipient cells [13,25,26],

  5. synthesizing from oligos reduced genomes designed to lack dispensable genes, and

  6. transplanting these reduced genomes into related Mycoplasma cells.

JCVI has completed steps (i)–(iv). Step (iv) was particularly challenging because of slow growth rates and because bacterial genomes engineered in yeast have DNA restriction/modification systems that are incompatible with the Mycoplasma host cell [26]. To learn how to transplant and express chemically synthesized genomes (iv), JCVI “booted-up” a synthetic, essentially wild-type, computer-specified, Mycoplasma mycoides genome (1,080 kbp) in a closely related cell to yield “Synthia” [13]. This technological milestone marks the dawn of “synthetic genomics” and will undoubtedly accelerate the engineering of microbial factories, once costs are significantly lowered, producing fuels, pharmaceuticals, chemicals, and novel biomaterials (see Prospects for Biotechnology). Notwithstanding the importance of this achievement, it should not be overinterpreted as synthesis of a cell or life, as standard usage of “synthetic” would imply either cell-free synthesis of the whole cell (rather than its genome) or generation of something very unnatural (rather than a genetically-modified organism). The published plan for steps (v) and (vi) is to synthesize a M. genitalium-based genome lacking all dispensable genes to boot up a “Mycoplasma laboratorium” cell (last paragraph of ref. [20]). However, though virtually all genes that are individually dispensable in M. genitalium have been determined, it is recognised that a major hurdle is synthetic lethals (i.e. non-viable cells when two individually viable mutations are combined [20]).

How may cellular complexity and synthetic lethality be circumvented to allow top-down production of a minimal genome? One route is step-wise deletion of the 100 individually dispensable genes, perhaps aided by directed evolution [24]. However, the number of combinations is astronomical, rational choice of combinations is limited by poor understanding (e.g., the functions of one fifth of the genes of M. genitalium remain to be determined), and considerably less than 100 of the 525 genes are likely dispensable in combination. There will also be multiple different minimal genome “solutions,” depending on the temporal order of deletion. Nevertheless, this will teach us much about redundancy in biology.

A second route is evident from tables of M. genitalium genes involved in the core replicative functions of DNA, RNA and protein syntheses [4]: these genes are in the minority, with the majority of M. genitalium genes being involved in functions such as metabolism of small molecules. Thus, if additional nutrients were supplied in the extracellular medium (and perhaps their uptake aided by encoding extra transmembrane transporters) it may be feasible to delete many more genes. This could take us down to a truly minimal, protein-coding cell: one sufficient for replication but not for metabolism of most small molecules.

Interestingly, development of such extreme metabolic dependence without loss of genetic independence may have already occurred in the reductive evolution of the intracellullar bacterial endosymbionts of insects [27]. These recently-sequenced symbiont genomes include the smallest non-organellar, non-viral genomes, Carsonella ruddii (160 kbp; 213 genes [28]) and Hodgkinia cicadicola (144 kbp; 188 genes [29]). In contrast to mitochondrial and chloroplast evolution, there is no evidence so far of gene transfer from bacterial symbiont to host [27]. Almost all of the core replicative functions have been predicted computationally to reside in the symbiont genome, although notable exceptions are several essential tRNAs and aminoacyl-tRNA synthetases [27,30]. Ultimate proof of genetic independence can only come from development of a defined in vitro system for replication of either a bacterial symbiont or a derivative engineered to encode any missing essential genes. Such experimental verification would constitute our third envisioned top-down route to a minimal genome.

As simple as these minimal cells may seem, it is worth noting that “there is no such thing as a simple bacterium” [31]. Mycoplasma pneumonia (only 816 kbp and 733 predicted genes) was recently found to have an unanticipated complexity that is humbling. Many genes have multiple modes of transcription and complicated regulation [32], the proteome has a similar organization to more complex organisms [33], and even metabolic enzymes perform multiple functions [34]. Furthermore, there is no rapid or systematic method for determining the functions of the large numbers of genes of unknown function in any organism, minimal or otherwise.

Bottom-up approach: in vitro construction

The alternative direction to a minimal cell is bottom-up: synthesizing self-replication by pooling together essential purified biological macromolecules, their genes and their small molecule substrates [4]. By this approach, cellular overhead including genes of unknown function can be removed, the system can be readily manipulated and tuned, and all of the components can be defined. One possibility is a DNA/RNA/protein system derived from the core replication machinery of today’s simplest cells. The other possibilities are ribonucleoprotein and RNA-alone systems modeling cells presumed to have existed billions of years ago [35].

Modeling the RNA world

A self-replicating system made solely from RNA [36] has the advantage of avoiding altogether the complexity of protein synthesis. Indeed, the milestone of self-sustained replication of an RNA enzyme in the absence of protein was just reached using pre-synthesized half-enzymes as substrates for ligation [37]. But this system cannot synthesize the half-enzyme substrates which are huge compared with natural small-molecule substrates and which contain all the informational content of the replicating system. A ribozyme selected from random sequences to polymerize nucleoside triphosphates on an RNA template was published 14 years ago [38] and its 3-dimensional structure just solved [39]. Yet the difficulty in developing this polymerase capable of adding only 14 nucleotides indicates that evolving it or random sequences in vitro into an RNA replicase is distant.

A protein-based in vitro minimal cell project (MCP)

A protein-based self-replicating system has the advantage of connecting with our current biological systems. Detailed plans to construct protein-based self-replication from small molecule substrates by combining already-reconstituted, purified, biochemical processes for DNA/RNA/protein syntheses [4] are essentially unchanged and under way. The proposal is to:

  1. identify the necessary genes,

  2. prepare efficient purified biochemical subsystems from the gene products,

  3. integrate the subsystems for self-replication (Fig. 1) and

  4. encapsulate the system within a membrane to give a synthetic cell (“synthetic life”).

Figure 1. Biochemical subsystems proposed to be sufficient for self-replication from supplied small molecule substrates.

Figure 1

Bold arrows indicate steps that have been largely integrated. The figure is adapted from Figs. 1 and 2 of ref. [4].

Of all the macromolecular components from E. coli and its bacteriophages, only 151 were hypothesized to be sufficient for the MCP, constituting a minigenome of 113 kbp [4]. Of these 151, it is striking that 96% are for protein synthesis and that there is considerable similarity in gene number and content and genome size to the recently-sequenced, extremely-metabolic-dependent, bacterial endosymbionts of insects (see above). An RNA/protein-based transcription/translation system has been reconstituted from purified components [40], but the omission of DNA does not simplify the number of genes that ultimately will be necessary to encode the whole system for self-replication. Rather, it creates a new set of challenges unsolved in the modern world: production of a functional large RNA genome that avoids inhibitory double-stranded RNA structures and replicative mutations [35].

Progress in step (i) has been rapid for E. coli (but slow for M. genitalium [41]). Of the missing 1–4 key ribosomal RNA (rRNA) modification genes, 3 have just been discovered [4244]. The gene for modifying transfer RNA (tRNA) A37 to t6A has also been found and shown to be essential for E. coli viability [45]. This only leaves as little as one other gene to find, involved in modifying tRNA U34 to cmo5U, with 2 genes in that pathway already known [46]. Thus, reconstitution from purified components of every subsystem of the MCP is tantalizingly near. In an attempt to close perhaps the biggest remaining gap, we are over-expressing the 5 known key rRNA modification enzymes [4] to test for activation of unmodified 23S rRNA transcripts necessary for synthesis of ribosomes in vitro.

Less progress has been reported on steps (ii)–(iv). With regard to step (ii), though the E. coli translation apparatus and ribosome were reconstituted separately from purified cellular components 3 decades ago, their translational accuracy is poorly characterized and in vitro efficiencies of protein synthesis and ribosome turnover remain low in both purified and crude systems (Table 1). The break-even milestone for ribosomes making all of the proteins in the proposed minigenome [4] is synthesis of ~35,000 peptide bonds by each ribosome (including 7491 peptide bonds for the ribosomal proteins). Towards the integration required for steps (iii) and (iv), bacterial transcription initiation has been reconstituted in a purified translation system [47], purified DNA-dependent transcription and translation has been performed within liposomes [48], and membrane proteins involved in phospholipid synthesis have been synthesized in active form in liposomes [49]. But some of the other subsystems require unphysiological conditions that preclude integration. Simple systems for DNA replication require thermocycling and oligo primers (PCR or circle-to-circle amplification [50]), while self-assembly of the E. coli ribosome from natural components requires low and high Mg2+ concentrations, high temperatures and long incubation times [51]. Nevertheless, physiological conditions for E. coli ribosome assembly have now been found and rRNA synthesis, ribosome assembly and translation (Fig. 1) have been integrated under batch conditions (Jewett and Church, submitted). The next steps will be substitution of the E. coli cells and extracts used for the macromolecule syntheses by purified subsystems.

Table 1.

Protein yields and costs in cell-free, transcription and translation systems from E. coli.

E. coli translation system Energy substrates and cofactors Reactor type Time (h) Protein product Protein yield (mg/mL) Protein yielda (AAs/ribosome) Protein costb ($/mg) Reference

S30 extract PEP, NAD, CoA, NTPs batch 3 CAT 0.75 6,600 16c [61]
S30 extract glucose, phosphate, NAD, CoA, NMPs batch 3 CAT 0.68 6,000 15c [66]
S30 extract glutamate, phosphate, NAD, CoA, NMPs, O2 batch 2 CAT 1.2 10,560 9c [60]
S30 extract PEP, proprietary mix (Roche RTS system) continuous exchange 100 GFP 1.00 8,400 370d [67]
S30 extract, condensed CrP, NTPs continuous exchange 21 CAT 6.00 26,000 6c [68]
PURE CrP, NTPs batch 3 GFP 0.30 2,700 2,900d New England Biolabs
a

For comparison, E. coli makes ~55,000 peptide bonds by each ribosome per cell doubling in 20 minutes. Concentrations of active ribosomes were assumed to be 2 μM for ref. [68] and 1 μM for the other systems (assuming ~50% ribosomes translating).

b

For comparison, in vivo production using E. coli can be as low as $0.005/mg [69].

c

Estimated cost for labor, equipment, consumables and reagents.

d

Based on kit price.

Abbreviations: AA, amino acids; PEP, phosphoenolpyruvate; CrP, creatine phosphate; CAT, chloramphenicol acetyl transferase; GFP, green fluorescent protein.

How might the efficiencies and utilities of purified systems be improved? There are some recent indications that adding genes not on the minimal list [4] should help. Inclusion of translation elongation factors not present in PURE kits might improve efficiency and/or accuracy: EF-P facilitates formation of the first peptide bond by positioning fMet-tRNAifMet [52], and LepA promotes back translocation of the mRNA-tRNA complex [53,54]. Comprehensive analysis of the individual effects of every E. coli protein on purified translation showed that 344 (8%) were stimulatory [55]. Most beneficial were ATP-dependent RNA helicase, HrpA, and trigger factor, increasing yields by ~80% and ~30%, respectively. More than 20 different auxiliary factors are thought to facilitate ribosome assembly, including chaperones, GTPases and helicases [56]. For example, ATP-dependent RNA helicase, DbpA, has specificity for 23S rRNA [57], and RimJ functions in ribosomal protein acetylation and in 30S subunit assembly [58]. Choices for gene addition will be informed by studies such as the measurement of kinetic effects on 30S assembly of Era, RimM and RimP [59]. Also, cytoplasmic mimicry has been shown to be a powerful guiding principle. Mimicking combined energy metabolism, oxidative phosphorylation and protein synthesis in crude extracts increased protein synthesis yields (Table 1; [60,61]). Activating natural energy metabolism in crude extracts reduces costs and suggests that incorporating metabolic modules [62] into the MCP could further increase utility.

It should be emphasized that genes other than the 151 may ultimately prove necessary for self-replication and that, while the MCP would certainly be helpful in revealing their existence, such mystery genes would be hard to identify. Identification may proceed through traditional biochemical purifications from extracts or by modern high throughput genetic screens [55]. Another challenge looming is how to achieve coordinated control of so many genes [18].

Prospects for biotechnology

Minimal cell syntheses are still in their formative stages where the main rewards are new molecular tools and a better understanding of the core genetic and biochemical systems necessary for basic life. But applications in biotechnology are close at hand. Based on the improved stability, growth and protein production of E. coli and other biotechnology workhorses upon reducing their genomes [2123], further minimized strains should replace most current commercial bacterial strains. Biotech applications of reduced-genome M. genitalium are less clear because of its fragility and much slower growth rate (doubling time in culture of 12 h). However, M. genitalium has the advantage of having the smallest genome, facilitating synthesis of variant genomes, and it is conceivable that its limitations might be addressed by synthetic genomics. Synthetic genomics will be particularly helpful for redesigning microbes for which genetic tools are poor.

The MCP mostly involves synthesis and optimization of purified translation systems. Such systems have a number of advantages over alternative methods of protein synthesis such as lack of RNases/proteases/inclusion bodies, high compatibility with cytotoxic proteins, flexibility of incorporation of unnatural amino acids, ease of product purification, and direct control of reaction conditions. The main hurdle preventing application of the PURE system in biotechnology is the high cost (Table 1) due to its production from >30 different fermentations. To address this limitation, we are developing a cost-effective method for over-expressing the entire system in a single E. coli cell followed by single batch purification [18,19,63]. Selection of variant 23S rRNAs for improved unnatural amino acid incorporation [64] could be uncoupled from cell viability by synthesizing ribosomes in vitro; such variants would facilitate the directed evolution of peptidomimetic drug candidates [65].

In conclusion, significant progress has been made in both the top-down and bottom-up approaches to minimal cells in the last 5 years. Both approaches are providing new tools, fundamental biological knowledge and potential biotech applications distinct from those garnered from other fields. Though major challenges lie ahead, the era of biology by design has begun.

Acknowledgments

We are grateful to George Church for advice and comments on the manuscript and John Glass, John McCutcheon and Michael Sismour for comments on the manuscript. This work was supported by the National Institutes of Health and National Academies Keck Futures Initiative (to MCJ and ACF), the National Science Foundation (to MCJ), and the American Cancer Society (to ACF).

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References and annotation

• Of special interest

•• Of outstanding interest

  • 1.Purnick PE, Weiss R. The second wave of synthetic biology: from modules to systems. Nat Rev Mol Cell Biol. 2009;10:410–422. doi: 10.1038/nrm2698. [DOI] [PubMed] [Google Scholar]
  • 2.Steen EJ, Kang Y, Bokinsky G, Hu Z, Schirmer A, McClure A, Del Cardayre SB, Keasling JD. Microbial production of fatty-acid-derived fuels and chemicals from plant biomass. Nature. 2010;463:559–562. doi: 10.1038/nature08721. [DOI] [PubMed] [Google Scholar]
  • 3.Tan C, Marguet P, You L. Emergent bistability by a growth-modulating positive feedback circuit. Nat Chem Biol. 2009;5:842–848. doi: 10.1038/nchembio.218. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 4.Forster AC, Church GM. Towards synthesis of a minimal cell. Mol Syst Biol. 2006;2:45. doi: 10.1038/msb4100090. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 5.Forster AC, Church GM. Synthetic biology projects in vitro. Genome Res. 2007;17:1–6. doi: 10.1101/gr.5776007. [DOI] [PubMed] [Google Scholar]
  • 6.Chiarabelli C, Stano P, Luisi PL. Chemical approaches to synthetic biology. Curr Opin Biotechnol. 2009;20:492–497. doi: 10.1016/j.copbio.2009.08.004. [DOI] [PubMed] [Google Scholar]
  • 7.Murtas G. Artificial assembly of a minimal cell. Mol Biosyst. 2009;5:1292–1297. doi: 10.1039/b906541e. [DOI] [PubMed] [Google Scholar]
  • 8.Fraser CM, Gocayne JD, White O, Adams MD, Clayton RA, Fleischmann RD, Bult CJ, Kerlavage AR, Sutton G, Kelley JM, et al. The minimal gene complement of Mycoplasma genitalium. Science. 1995;270:397–403. doi: 10.1126/science.270.5235.397. [DOI] [PubMed] [Google Scholar]
  • 9.Gibson DG, Benders GA, Andrews-Pfannkoch C, Denisova EA, Baden-Tillson H, Zaveri J, Stockwell TB, Brownley A, Thomas DW, Algire MA, et al. Complete chemical synthesis, assembly, and cloning of a Mycoplasma genitalium genome. Science. 2008;319:1215–1220. doi: 10.1126/science.1151721. [DOI] [PubMed] [Google Scholar]
  • 10.Gibson DG, Young L, Chuang RY, Venter JC, Hutchison CA, 3rd, Smith HO. Enzymatic assembly of DNA molecules up to several hundred kilobases. Nat Methods. 2009;6:343–345. doi: 10.1038/nmeth.1318. [DOI] [PubMed] [Google Scholar]
  • 11•.Gibson DG, Benders GA, Axelrod KC, Zaveri J, Algire MA, Moodie M, Montague MG, Venter JC, Smith HO, Hutchison CA., 3rd One-step assembly in yeast of 25 overlapping DNA fragments to form a complete synthetic Mycoplasma genitalium genome. Proc Natl Acad Sci U S A. 2008;105:20404–20409. doi: 10.1073/pnas.0811011106. This surprising property of yeast simplified construction and manipulation of synthetic genomes. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 12.Benders GA, Noskov VN, Denisova EA, Lartigue C, Gibson DG, Assad-Garcia N, Chuang RY, Carrera W, Moodie M, Algire MA, et al. Cloning whole bacterial genomes in yeast. Nucleic Acids Res. 2010 doi: 10.1093/nar/gkq119. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 13••.Gibson DG, Glass JI, Lartigue C, Noskov VN, Chuang RY, Algire MA, Benders GA, Montague MG, Ma L, Moodie MM, et al. Creation of a Bacterial Cell Controlled by a Chemically Synthesized Genome. Science. 2010 doi: 10.1126/science.1190719. This major technological milestone demonstrated that entire synthesized genomes can be implanted and booted-up in cells. [DOI] [PubMed] [Google Scholar]
  • 14.Gibson DG. Synthesis of DNA fragments in yeast by one-step assembly of overlapping oligonucleotides. Nucleic Acids Res. 2009;37:6984–6990. doi: 10.1093/nar/gkp687. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 15.Tian J, Gong H, Sheng N, Zhou X, Gulari E, Gao X, Church G. Accurate multiplex gene synthesis from programmable DNA microchips. Nature. 2004;432:1050–1054. doi: 10.1038/nature03151. [DOI] [PubMed] [Google Scholar]
  • 16.Carr PA, Church GM. Genome engineering. Nat Biotechnol. 2009;27:1151–1162. doi: 10.1038/nbt.1590. [DOI] [PubMed] [Google Scholar]
  • 17.Kwok R. Five hard truths for synthetic biology. Nature. 2010;463:288–290. doi: 10.1038/463288a. [DOI] [PubMed] [Google Scholar]
  • 18.Du L, Gao R, Forster AC. Engineering multigene expression in vitro and in vivo with small terminators for T7 RNA polymerase. Biotechnol Bioeng. 2009;104:1189–1196. doi: 10.1002/bit.22491. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 19••.Wang HH, Isaacs FJ, Carr PA, Sun ZZ, Xu G, Forest CR, Church GM. Programming cells by multiplex genome engineering and accelerated evolution. Nature. 2009;460:894–898. doi: 10.1038/nature08187. This invention enables rapid cell mutagenesis for metabolic engineering and synthetic biology. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 20.Glass JI, Assad-Garcia N, Alperovich N, Yooseph S, Lewis MR, Maruf M, Hutchison CA, 3rd, Smith HO, Venter JC. Essential genes of a minimal bacterium. Proc Natl Acad Sci U S A. 2006;103:425–430. doi: 10.1073/pnas.0510013103. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 21••.Posfai G, Plunkett G, 3rd, Feher T, Frisch D, Keil GM, Umenhoffer K, Kolisnychenko V, Stahl B, Sharma SS, de Arruda M, et al. Emergent properties of reduced-genome Escherichia coli. Science. 2006;312:1044–1046. doi: 10.1126/science.1126439. Surprisingly, targeted deletions removing 15% of E. coli’s genome were not only viable, they improved its properties for applications in molecular biology. [DOI] [PubMed] [Google Scholar]
  • 22.Mizoguchi H, Sawano Y, Kato J, Mori H. Superpositioning of deletions promotes growth of Escherichia coli with a reduced genome. DNA Res. 2008;15:277–284. doi: 10.1093/dnares/dsn019. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 23.Morimoto T, Kadoya R, Endo K, Tohata M, Sawada K, Liu S, Ozawa T, Kodama T, Kakeshita H, Kageyama Y, et al. Enhanced recombinant protein productivity by genome reduction in Bacillus subtilis. DNA Res. 2008;15:73–81. doi: 10.1093/dnares/dsn002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 24.Nilsson AI, Koskiniemi S, Eriksson S, Kugelberg E, Hinton JC, Andersson DI. Bacterial genome size reduction by experimental evolution. Proc Natl Acad Sci U S A. 2005;102:12112–12116. doi: 10.1073/pnas.0503654102. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 25•.Lartigue C, Glass JI, Alperovich N, Pieper R, Parmar PP, Hutchison CA, 3rd, Smith HO, Venter JC. Genome transplantation in bacteria: changing one species to another. Science. 2007;317:632–638. doi: 10.1126/science.1144622. This demonstrated that the genome of one bacterium can be transplanted into a different bacterium and transform the cell, setting the stage for synthetic genomics. [DOI] [PubMed] [Google Scholar]
  • 26.Lartigue C, Vashee S, Algire MA, Chuang RY, Benders GA, Ma L, Noskov VN, Denisova EA, Gibson DG, Assad-Garcia N, et al. Creating bacterial strains from genomes that have been cloned and engineered in yeast. Science. 2009;325:1693–1696. doi: 10.1126/science.1173759. [DOI] [PubMed] [Google Scholar]
  • 27.McCutcheon JP. The bacterial essence of tiny symbiont genomes. Curr Opin Microbiol. 2010;13:73–78. doi: 10.1016/j.mib.2009.12.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 28••.Nakabachi A, Yamashita A, Toh H, Ishikawa H, Dunbar HE, Moran NA, Hattori M. The 160-kilobase genome of the bacterial endosymbiont Carsonella. Science. 2006;314:267. doi: 10.1126/science.1134196. This genome sequence from a new class of non-viral, non-organelle genomes was the first found to be much smaller than that of M. genitalium. [DOI] [PubMed] [Google Scholar]
  • 29•.McCutcheon JP, McDonald BR, Moran NA. Origin of an alternative genetic code in the extremely small and GC-rich genome of a bacterial symbiont. PLoS Genet. 2009;5:e1000565. doi: 10.1371/journal.pgen.1000565. This is the smallest non-viral, non-organelle genome sequence. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 30.McCutcheon JP, McDonald BR, Moran NA. Convergent evolution of metabolic roles in bacterial co-symbionts of insects. Proc Natl Acad Sci U S A. 2009;106:15394–15399. doi: 10.1073/pnas.0906424106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 31.Ochman H, Raghavan R. Systems biology. Excavating the functional landscape of bacterial cells. Science. 2009;326:1200–1201. doi: 10.1126/science.1183757. [DOI] [PubMed] [Google Scholar]
  • 32.Guell M, van Noort V, Yus E, Chen WH, Leigh-Bell J, Michalodimitrakis K, Yamada T, Arumugam M, Doerks T, Kuhner S, et al. Transcriptome complexity in a genome-reduced bacterium. Science. 2009;326:1268–1271. doi: 10.1126/science.1176951. [DOI] [PubMed] [Google Scholar]
  • 33.Kuhner S, van Noort V, Betts MJ, Leo-Macias A, Batisse C, Rode M, Yamada T, Maier T, Bader S, Beltran-Alvarez P, et al. Proteome organization in a genome-reduced bacterium. Science. 2009;326:1235–1240. doi: 10.1126/science.1176343. [DOI] [PubMed] [Google Scholar]
  • 34.Yus E, Maier T, Michalodimitrakis K, van Noort V, Yamada T, Chen WH, Wodke JA, Guell M, Martinez S, Bourgeois R, et al. Impact of genome reduction on bacterial metabolism and its regulation. Science. 2009;326:1263–1268. doi: 10.1126/science.1177263. [DOI] [PubMed] [Google Scholar]
  • 35.Poole AM, Logan DT. Modern mRNA proofreading and repair: clues that the last universal common ancestor possessed an RNA genome? Mol Biol Evol. 2005;22:1444–1455. doi: 10.1093/molbev/msi132. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 36.Doudna JA, Szostak JW. RNA-catalysed synthesis of complementary-strand RNA. Nature. 1989;339:519–522. doi: 10.1038/339519a0. [DOI] [PubMed] [Google Scholar]
  • 37.Lincoln TA, Joyce GF. Self-sustained replication of an RNA enzyme. Science. 2009;323:1229–1232. doi: 10.1126/science.1167856. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 38.Ekland EH, Bartel DP. RNA-catalysed RNA polymerization using nucleoside triphosphates. Nature. 1996;382:373–376. doi: 10.1038/382373a0. [DOI] [PubMed] [Google Scholar]
  • 39.Shechner DM, Grant RA, Bagby SC, Koldobskaya Y, Piccirilli JA, Bartel DP. Crystal structure of the catalytic core of an RNA-polymerase ribozyme. Science. 2009;326:1271–1275. doi: 10.1126/science.1174676. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 40.Kita H, Matsuura T, Sunami T, Hosoda K, Ichihashi N, Tsukada K, Urabe I, Yomo T. Replication of genetic information with self-encoded replicase in liposomes. Chembiochem. 2008;9:2403–2410. doi: 10.1002/cbic.200800360. [DOI] [PubMed] [Google Scholar]
  • 41.de Crecy-Lagard V, Marck C, Brochier-Armanet C, Grosjean H. Comparative RNomics and modomics in Mollicutes: prediction of gene function and evolutionary implications. IUBMB Life. 2007;59:634–658. doi: 10.1080/15216540701604632. [DOI] [PubMed] [Google Scholar]
  • 42•.Purta E, O’Connor M, Bujnicki JM, Douthwaite S. YgdE is the 2′-O-ribose methyltransferase RlmM specific for nucleotide C2498 in bacterial 23S rRNA. Mol Microbiol. 2009;72:1147–1158. doi: 10.1111/j.1365-2958.2009.06709.x. This identifies an E. coli rRNA modification enzyme and tabulates all such enzymes. [DOI] [PubMed] [Google Scholar]
  • 43.Lesnyak DV, Sergiev PV, Bogdanov AA, Dontsova OA. Identification of Escherichia coli m2G methyltransferases: I. the ycbY gene encodes a methyltransferase specific for G2445 of the 23 S rRNA. J Mol Biol. 2006;364:20–25. doi: 10.1016/j.jmb.2006.09.009. [DOI] [PubMed] [Google Scholar]
  • 44.Toh SM, Xiong L, Bae T, Mankin AS. The methyltransferase YfgB/RlmN is responsible for modification of adenosine 2503 in 23S rRNA. RNA. 2008;14:98–106. doi: 10.1261/rna.814408. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 45.El Yacoubi B, Lyons B, Cruz Y, Reddy R, Nordin B, Agnelli F, Williamson JR, Schimmel P, Swairjo MA, de Crecy-Lagard V. The universal YrdC/Sua5 family is required for the formation of threonylcarbamoyladenosine in tRNA. Nucleic Acids Res. 2009;37:2894–2909. doi: 10.1093/nar/gkp152. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 46.Nasvall SJ, Chen P, Bjork GR. The modified wobble nucleoside uridine-5-oxyacetic acid in tRNAPro(cmo5UGG) promotes reading of all four proline codons in vivo. RNA. 2004;10:1662–1673. doi: 10.1261/rna.7106404. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 47.Asahara H, Chong S. In vitro genetic reconstruction of bacterial transcription initiation by coupled synthesis and detection of RNA polymerase holoenzyme. Nucleic Acids Res. 2010 doi: 10.1093/nar/gkq377. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 48.Murtas G, Kuruma Y, Bianchini P, Diaspro A, Luisi PL. Protein synthesis in liposomes with a minimal set of enzymes. Biochem Biophys Res Commun. 2007;363:12–17. doi: 10.1016/j.bbrc.2007.07.201. [DOI] [PubMed] [Google Scholar]
  • 49.Kuruma Y, Stano P, Ueda T, Luisi PL. A synthetic biology approach to the construction of membrane proteins in semi-synthetic minimal cells. Biochim Biophys Acta. 2009;1788:567–574. doi: 10.1016/j.bbamem.2008.10.017. [DOI] [PubMed] [Google Scholar]
  • 50.Dahl F, Baner J, Gullberg M, Mendel-Hartvig M, Landegren U, Nilsson M. Circle-to-circle amplification for precise and sensitive DNA analysis. Proc Natl Acad Sci U S A. 2004;101:4548–4553. doi: 10.1073/pnas.0400834101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 51.Nierhaus KH. In: Reconstitution of ribosomes, in Ribosomes and Protein Synthesis, A Practical Approach. Spedding G, editor. Oxford: Oxford University Press; 1990. [Google Scholar]
  • 52.Blaha G, Stanley RE, Steitz TA. Formation of the first peptide bond: the structure of EF-P bound to the 70S ribosome. Science. 2009;325:966–970. doi: 10.1126/science.1175800. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 53•.Qin Y, Polacek N, Vesper O, Staub E, Einfeldt E, Wilson DN, Nierhaus KH. The highly conserved LepA is a ribosomal elongation factor that back-translocates the ribosome. Cell. 2006;127:721–733. doi: 10.1016/j.cell.2006.09.037. This reports demonstrates that LepA is an elongation factor that enables a unique back-translocation function during protein synthesis. [DOI] [PubMed] [Google Scholar]
  • 54.Evans RN, Blaha G, Bailey S, Steitz TA. The structure of LepA, the ribosomal back translocase. Proc Natl Acad Sci U S A. 2008;105:4673–4678. doi: 10.1073/pnas.0801308105. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 55•.Kazuta Y, Adachi J, Matsuura T, Ono N, Mori H, Yomo T. Comprehensive analysis of the effects of Escherichia coli ORFs on protein translation reaction. Mol Cell Proteomics. 2008;7:1530–1540. doi: 10.1074/mcp.M800051-MCP200. This is a high throughput genetic screen testing the effect of every E. coli protein individually on purified translation. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 56.Wilson DN, Nierhaus KH. The weird and wonderful world of bacterial ribosome regulation. Crit Rev Biochem Mol Biol. 2007;42:187–219. doi: 10.1080/10409230701360843. [DOI] [PubMed] [Google Scholar]
  • 57.Sharpe Elles LM, Sykes MT, Williamson JR, Uhlenbeck OC. A dominant negative mutant of the E. coli RNA helicase DbpA blocks assembly of the 50S ribosomal subunit. Nucleic Acids Res. 2009;37:6503–6514. doi: 10.1093/nar/gkp711. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 58.Roy-Chaudhuri B, Kirthi N, Kelley T, Culver GM. Suppression of a cold-sensitive mutation in ribosomal protein S5 reveals a role for RimJ in ribosome biogenesis. Mol Microbiol. 2008;68:1547–1559. doi: 10.1111/j.1365-2958.2008.06252.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 59.Bunner AE, Nord S, Wikstrom PM, Williamson JR. The Effect of Ribosome Assembly Cofactors on In Vitro 30S Subunit Reconstitution. J Mol Biol. 2010;398:1–7. doi: 10.1016/j.jmb.2010.02.036. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 60.Jewett MC, Calhoun KA, Voloshin A, Wuu JJ, Swartz JR. An integrated cell-free metabolic platform for protein production and synthetic biology. Mol Syst Biol. 2008;4:220. doi: 10.1038/msb.2008.57. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 61.Jewett MC, Swartz JR. Mimicking the Escherichia coli cytoplasmic environment activates long-lived and efficient cell-free protein synthesis. Biotechnol Bioeng. 2004;86:19–26. doi: 10.1002/bit.20026. [DOI] [PubMed] [Google Scholar]
  • 62.Bujara M, Schumperli M, Billerbeck S, Heinemann M, Panke S. Exploiting cell free systems: Implementation and debugging of a system of biotransformations. Biotechnol Bioeng. 2010;106:376–389. doi: 10.1002/bit.22666. [DOI] [PubMed] [Google Scholar]
  • 63.Ederth J, Mandava CS, Dasgupta S, Sanyal S. A single-step method for purification of active His-tagged ribosomes from a genetically engineered Escherichia coli. Nucleic Acids Res. 2009;37:e15. doi: 10.1093/nar/gkn992. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 64.Dedkova LM, Fahmi NE, Golovine SY, Hecht SM. Enhanced D-amino acid incorporation into protein by modified ribosomes. J Am Chem Soc. 2003;125:6616–6617. doi: 10.1021/ja035141q. [DOI] [PubMed] [Google Scholar]
  • 65.Tan Z, Blacklow SC, Cornish VW, Forster AC. De novo genetic codes and pure translation display. Methods. 2005;36:279–290. doi: 10.1016/j.ymeth.2005.04.011. [DOI] [PubMed] [Google Scholar]
  • 66.Calhoun KA, Swartz JR. An economical method for cell-free protein synthesis using glucose and nucleoside monophosphates. Biotechnol Prog. 2005;21:1146–1153. doi: 10.1021/bp050052y. [DOI] [PubMed] [Google Scholar]
  • 67.Noireaux V, Libchaber A. A vesicle bioreactor as a step toward an artificial cell assembly. Proc Natl Acad Sci U S A. 2004;101:17669–17674. doi: 10.1073/pnas.0408236101. [DOI] [PMC free article] [PubMed] [Google Scholar]
  • 68.Kigawa T, Yabuki T, Yoshida Y, Tsutsui M, Ito Y, Shibata T, Yokoyama S. Cell-free production and stable-isotope labeling of milligram quantities of proteins. FEBS Lett. 1999;442:15–19. doi: 10.1016/s0014-5793(98)01620-2. [DOI] [PubMed] [Google Scholar]
  • 69.Swartz JR. Advances in Escherichia coli production of therapeutic proteins. Curr Opin Biotechnol. 2001;12:195–201. doi: 10.1016/s0958-1669(00)00199-3. [DOI] [PubMed] [Google Scholar]

RESOURCES