Abstract
Cell-free gene expression (CFE) systems empower synthetic biologists to build biological molecules and processes outside of living intact cells. The foundational principle is that precise, complex biomolecular transformations can be conducted in purified enzyme or crude cell lysate systems. This concept circumvents mechanisms that have evolved to facilitate species survival, bypasses limitations on molecular transport across the cell wall, and provides a significant departure from traditional, cell-based processes that rely on microscopic cellular “reactors.” In addition, cell-free systems are inherently distributable through freeze-drying, which allows simple distribution before rehydration at the point-of-use. Furthermore, as cell-free systems are nonliving, they provide built-in safeguards for biocontainment without the constraints attendant on genetically modified organisms. These features have led to a significant increase in the development and use of CFE systems over the past two decades. Here, we discuss recent advances in CFE systems and highlight how they are transforming efforts to build cells, control genetic networks, and manufacture biobased products.
1. Introduction
Cell-free biology, the activation of biological processes without the use of intact living cells,1 provides a versatile approach for understanding, building, and harnessing biological systems. Cell-free biology offers several advantages over living systems by bypassing cell growth and viability constraints, enabling customization of biosynthesis conditions for a single product, and offering unprecedented flexibility in designing and controlling biological systems. The scope of the term “cell-free systems” is broad, as it has been applied to many biological samples or mixtures that do not contain whole cells. Here we focus on cell-free gene expression (CFE), the process of producing proteins using transcription and translation machinery extracted from cells and carried out in a controlled, nonliving environment.
Cell-free studies have played a crucial role in the elucidation of biological mechanisms since the late 19th and early 20th century.2 These foundational investigations include the observation of fermentation in extract from yeast cells,3,4 the subsequent reconstitution of glycolysis with purified enzymes,5 the determination of the cell cycle using eukaryotic cell extracts,6 and the discovery of the genetic code.7,8 The latter led to the development of CFE systems, which form the core of many modern cell-free applications by activating transcription and/or translation machinery in purified systems or crude cell extracts to execute biologically encoded genetic programs.
CFE was originally developed as a simplified method to study protein biosynthesis.2 From 1948 to 1960, several researchers used crude cell extracts (comprising the soluble components of lysed cells, often mixed with exogenous reagents) to assess the incorporation of radiolabeled amino acids into proteins, studying the mechanism of peptide bond formation without much knowledge of translation beyond the essentiality of ATP, GTP, and tRNAs.9−12 A turning point came in 1961, when Nirenberg and Matthaei uncovered the dependence of protein synthesis on RNA templates in extract from Escherichia coli.7 Several years of investigation uncovered the entire set of RNA triplets encoding each amino acid8 in addition to the canonical start and stop codons.13 This pioneering work enabled the translation of proteins in a wide variety of cell extracts beyond E. coli, including rabbit reticulocytes and wheat germ, but all initial systems relied on RNA templates with reactions lasting on the order of minutes.13
In the 1970s, Zubay demonstrated CFE from DNA templates,14 thereby recapitulating combined transcription and translation outside living cells, enabling longer lasting reactions and incorporating genetic regulatory elements into CFE; yet, these reactions produced only trace amounts of protein measured by radioactive counts. The limitations were addressed in part using viral RNA polymerases in the 1980s-90s,15,16 and with improvements in reaction conditions,17 metabolism activation,18,19 and extract preparation developed in the late 1990s and early 2000s.2,20 Modern E. coli CFE systems now routinely reach gram per liter yields of protein in batch reactions.21−24 Innovations in CFE systems derived from wheat germ cells also led to g/L yields as early as 2000,25 though in continuous exchange systems as we will discuss later. While most of these studies utilized reporter proteins such as chloramphenicol acetyltransferase or green fluorescent protein (GFP), the simultaneous optimization of CFE in E. coli extract and the development of a fully purified transcription and translation platform (PURE: Protein synthesis Using Recombinant Elements) in the early 2000s26 led to an explosion of research applications enabled by rapid protein synthesis in vitro. Common uses of CFE now include the production of toxic or post-translationally modified proteins, making medicines at the point-of-use, the incorporation of noncanonical amino acids into proteins, the screening of genetic parts and enzymes, and the generation of synthetic biological systems (e.g., artificial cells or engineered ribosomes).
The primary objective of this review is to provide a detailed roadmap of the ways in which CFE systems are manufactured and how they have been applied to diverse problems in science and engineering. This review focuses largely on CFE systems that involve both transcription and translation. There have been many other excellent reviews in this space,1,27−31 and we aimed to update and collate this information here. We start by providing an overview of the different types of CFE systems and the different components required for activity (Section 2). We focus on E. coli derived CFE systems as they are the most well-developed and widely used. We then detail the diverse applications of CFE in order of increasing biological complexity (Section 3), from the expression of single genes up to the construction of synthetic cells.
2. Methods for Cell-Free Gene Expression
Cell-free gene expression (CFE), also referred to as in vitro transcription and translation (IVTT or TX-TL)21,32 or cell-free protein synthesis (CFPS),1 is the activation of transcription and translation from cellular components derived from either crude cellular extracts of various organisms28 (e.g., E. coli, yeast, Chinese Hamster Ovary (CHO) cells, etc.) or from purified components as in the PURE system.26 CFE reactions are carried out by combining these components with a DNA (combined transcription and translation) or RNA (translation only) template and a cocktail of other molecules (e.g., NTPs, amino acids) required to activate and sustain transcription and translation.33 Run in batch mode, the highest yielding CFE reactions can reach up to 4 mg/mL21 and >20 h22 of protein synthesis, while semicontinuous reactions (i.e., semipermeable membrane exchange of small molecules) can reach yields of 8 mg/mL21 and 14 days34 of sustained protein synthesis. CFE reactions are scalable, with batch reaction volumes ranging from femtoliters in microfluidic systems35,36 to >100 L in traditional large-scale bioreactors.37 Trends in CFE systems over the past 20 years are summarized in Figure 1 and Table 1.
Figure 1.

Summary of key metrics for E. coli CFE over time. (A) Optimization of CFE using E. coli extract has led to high yields of several grams protein per liter of reaction in both batch and continuous modes. (B) Cost per mg of synthesized protein has decreased with time. Note that costs plotted here include only the the nucleoside tri- or mono-phosphates and carbon/energy sources. (C) The longevity of batch reactions has been extended to ∼20 h.
Table 1. Cost and Productivity of CFE Formulations over Timea.
| Year | Yield (mg/mL) | Type | Time (h) | Protein | Cost* ($/L Rxn) | $/mg Protein | Reference |
|---|---|---|---|---|---|---|---|
| 1996 | 0.4 | Batch | 1 | CAT | 6290 | 15.72 | Kim et al., 199638 |
| 2004 | 0.7 | Batch | 6 | CAT | 845 | 1.21 | Jewett and Swartz, 200418 |
| 2005 | 0.56 | Batch | 4 | CAT | 3716 | 6.64 | Voloshin and Swartz, 200539 |
| 2005 | 0.7 | Batch | 3 | CAT | 58 | 0.09 | Calhoun and Swartz, 200540 |
| 2008 | 0.74 | Batch | 3 | CAT | 3716 | 5.02 | Jewett et al., 200819 |
| 0.62 | Batch | 6 | CAT | 58 | 0.09 | ||
| 2011 | 0.7 | Batch | 10 | GM-CSF | 96 | 0.14 | Zawada et al., 201137 |
| 2011 | 1.7 | Batch | 12 | CAT | 811 | 0.48 | Kim et al., 201141 |
| 2012 | 0.7 | Batch | 8 | GFP | 3553 | 5.08 | Shin and Noireaux, 201242 |
| 2014 | 2.3 | Batch | 10 | GFP | 3555 | 1.55 | Caschera and Noireaux, 201424 |
| 2015 | 1.5 | Batch | 8 | Antibody | 75 | 0.05 | Cai et al., 201543 |
| 2016 | 2.03 | Batch | 16 | GFP | 3555 | 1.75 | Garamella et al., 201644 |
| 2018 | 1.78 | Batch | 20 | GFP | 4006 | 2.25 | Martin et al., 201822 |
| 2019 | 2.67 | Batch | 20 | GFP | 4006 | 1.50 | Des Soye et al., 201923 |
| 2021 | 4.05 | Batch | 20 | GFP | 3592 | 0.89 | Garenne et al., 202121 |
| 2023 | 1 | Batch | 20 | GFP | 90 | 0.09 | Warfel et al., 202345 |
| 1996 | 1.2 | CE | 14 | CAT | Kim and Choi, 199617 | ||
| 1999 | 6 | CE | 20 | CAT | Kigawa et al., 199946 | ||
| 2004 | 1 | CE | 100 | GFP | Noireaux and Libchaber, 200447 | ||
| 2005 | 6 | CE | 20 | GPCR | Klammt et al., 200548 | ||
| 2012 | 2.5 | CE | 20 | GFP | Shin and Noireaux, 201242 | ||
| 2016 | 6.35 | CE | 24 | GFP | Garamella et al., 201644 | ||
| 2021 | 8 | CE | 20 | GFP | Garenne et al., 202121 |
Selected publications from 1996-2023 highlight the trends in CFE metrics over decades of optimization in batch and continuous exchange (CE) formats. These data are graphically displayed in Figure 1. *Note that the reaction cost estimate includes only the nucleotides and carbon/energy sources, as the primary determinants of low-cost CFE are using NMPs and non-phosphorylated carbon substrates. Costs were not determined for continuous exchange reactions due to the large and variable volumes of feeding solution. Reporter proteins include chloramphenicol acetyltransferase (CAT), granulocyte-macrophage colony-stimulating factor (GM-CSF), green fluorescent protein (GFP), antibodies, and a G-protein coupled receptor (GPCR).
In the subsections that follow, we discuss extract-based CFE systems, extract preparation methods, CFE reaction components and energy systems, purified CFE systems, CFE reaction formats and scale, monitoring CFE reactions, high-throughput experimentation for CFE, and modeling CFE systems. In Section 3, we will cover applications.
2.1. Extract-Based CFE Systems
To produce proteins of interest, CFE systems harness an ensemble of catalytic components necessary for transcription, protein biosynthesis, protein folding, and energy regeneration from crude cell extracts. Crude cell extracts from a broad range of organisms have been employed. While early studies focused on E. coli,14 rabbit reticulocyte,49 and wheat germ systems,50 extracts from a more diverse array of organisms are now used to exploit differences in post-translational modifications, metabolism, temperature optima, and other properties.20,28 Here, we provide a brief overview of prokaryotic (Table 2) and eukaryotic (Table 3) CFE platforms.
Table 2. Example Prokaryotic CFE Systemsa.
| Source Organism | Synthesized Protein (μg/mL) | Reaction Format | Time (h) | Reference |
|---|---|---|---|---|
| Bacillus megaterium | 134 | Combined batch | 4 | Moore et al., 201851 |
| Bacillus subtilis | 21 | Combined batch | 2.5 | Kelwick et al., 201652 |
| 561 ± 8.7 | Combined batch | 13 | Zhang et al., 202153 | |
| Clostridium autoethanogeum | 236 ± 24 | Combined batch | 3 | Krüger et al., 202054 |
| Corynebacterium glutamicum | 385 ± 7 | Combined batch | 13 | Zhang et al., 202153 |
| Cutibacterium acnes | 85 | Combined batch | 24 | Fábrega et al., 202155 |
| Escherichia coli | 8000 | Combined continuous | 24 | Garenne et al., 202121 |
| 4050 | Combined batch | 20 | ||
| Escherichia fergusonii | 50 | Combined batch | 8 | Yim et al., 201956 |
| Klebsiella oxytoca, | 4 | |||
| Pantoea agglomerans | 150 | |||
| Pseudomonas fluorescens | 95 | Combined batch | 10 | Nakashima et al., 200457 |
| Pseudomonas putida | 198 ± 6 | Combined batch | 4 | Wang et al., 201858 |
| Streptomyces coelicolor | 30 | Combined batch | 3 | Li et al., 201759 |
| Streptomyces lividans | 516 ± 25 | Combined batch | 8 | Xu et al., 202260 |
| 400 | Combined continuous | 48 | Xu et al., 202061 | |
| Streptomyces venezuelae | 266 | Combined batch | 3 | Moore et al., 202162 |
| Thermococcus kodakaraensis | 115 | Batch translation | 0.25 | Endoh et al., 200763 |
| Vibrio natriegens | 1600 ± 50 | Combined batch | 20 | Des Soye et al., 201864 |
| Yersinia pestis | 30 | Combined batch | 16 | McDonald et al., 202165 |
Organisms are listed alphabetically. Synthesized protein is reported as mean ± standard deviation when error was enumerated in the original publication.
Table 3. Example Eukaryotic CFE Systemsa.
| Source Organism | Synthesized Protein (μg/mL) | Reaction Format | Time (h) | Reference |
|---|---|---|---|---|
| Bombyx mori | 70 | Batch translation | n.r. | Ito, 202166 |
| CHO | 982 ± 30 | Combined continuous | 48 | Thoring et al., 202167 |
| 50.2 | Batch | 2 | Brödel et al., 201768 | |
| HeLa | 50 | Continuous translation | 36 | Mikami et al., 200669 |
| Leishmania tarentolae | 300 | Batch translation | 1.5 | Mureev et al., 200970 |
| Nicotiana tabacum | 3000 | Combined batch | 48 | Das Gupta et al., 202371 |
| Pichia pastoris | 116 | Combined batch | 4 | Spice et al., 202072 |
| Rabbit reticulocyte | 40 | Batch translation | 1.5 | Mureev et al., 200970 |
| Saccharomyces cerevisiae | 20 ± 1.3 | Combined batch | 5 | Schoborg et al., 201673 |
| 17.0 ± 3.8 | Combined continuous | 10 | Schoborg et al., 201474 | |
| Spodoptera frugiperda | 40 | Combined batch | 1.5 | Merk et al., 201575 |
| 700 | Combined continuous | 72 | ||
| Trichoplusia ni | 71 | Batch translation | 6 | Ezure et al., 200676 |
| Wheat germ | 5000 | Continuous translation | 18 | Beebe et al., 201177 |
| 1600 | Batch | n.r. | Harbers, 201478 |
Organisms are listed alphabetically. Synthesized protein is reported as mean ± standard deviation when error was enumerated in the original publication. n.r. = not reported.
2.1.1. Prokaryotic Cell-Free Systems
E. coli cell-free systems are the most robust and well-characterized CFE platforms.1,30 Over the last 30 years, extract preparation and CFE protocols have been optimized by several groups to increase protein production to > g/L quantities (Figure 1, Table 2). These works have shown that requirements for maximal cell-free expression include adequate substrate supply, a homeostatic environment, and the removal or avoidance of inhibitory byproducts. These efforts have also taught us that crude extract cell-free systems should be thought of as a complex set of biochemical reactions rather than a “black box”. Current efforts are underway to understand the molecular composition of E. coli extracts and their impact on activity.79−83 For example, it is now known that proteomic shifts in extract preparations occur due to stress response,81 altered media formulations,84 engineered strains,85 as well as culture formats and lysis methods.86 Similarly, metabolomic analyses have characterized molecular changes in CFE reactions due to variations in extract source strain growth media and harvest conditions,87 the impact of metabolites on CFE,88 metabolic flux over time in differentially lysed extracts,89 and batch-to-batch variability.90
While E. coli extracts are used most predominantly, other bacterial CFE systems have recently emerged. These systems provide advantages for specific applications, such as genetic part prototyping or unique metabolism,51,54,56,58,91 but they have been less optimized and typically produce less protein than E. coli-based systems (Table 2). CFE systems from Vibrio natriegens are one exception. With little optimization, a V. natriegens-based CFE system was able to produce up to ∼400 μg/mL GFP from endogenous promoters92 and up to ∼1,600 μg/mL using T7 RNA polymerase64,93 with plasmid DNA templates. While bacterial CFE systems are routinely used for their protein biosynthesis activity and lower costs, they can lack chaperones and foldases that facilitate folding of eukaryotic proteins and in addition lack machinery for post-translational modifications. Such machinery can be added to the system, as we will discuss later in Section 3.1.
2.1.2. Eukaryotic Cell-Free Systems
Eukaryotic CFE systems generally have more laborious extract preparation procedures and are more costly than bacterial CFE systems due to slower growth rates, more specialized media components, and phosphorylated energy substrates. However, they can have advantages for expressing complex proteins, such as those with post-translational modifications. Rabbit reticulocyte and wheat germ based systems have been the most widely used eukaryotic CFE systems.28
Wheat germ extract is the most widespread CFE platform other than E. coli,28 and it has facilitated structural biology,94 synthetic biology,95 and pharmaceutical96 studies since its initial development for scientific research.78,97 This platform benefits from commercial availability, demonstrated scalability, and batch protein yields greater than 1 g/L.77,78
More recently, a high-yielding plant CFE platform derived from Nicotiana tabacum (tobacco) BY-2 cell cultures has been developed, which can be harvested and lysed much more rapidly than wheat germ.98 Tobacco BY-2 extract can express more than 3 g/L of protein71 and has successfully produced transmembrane, disulfide bonded, and glycosylated proteins,99 and has been scaled to 1 L reactions.100 Less optimized sources of translationally active plant extracts have come from other components of tobacco,101,102Arabidopsis thaliana callus,103Pisum sativum chloroplasts,104Zea mays seeds,105 and leaves from Z. mays and Beta vulgaris.106
Chinese hamster ovary (CHO) cells have also been used for eukaryotic CFE systems due to their abundance in commercial synthesis of biologics, such as antibodies. CHO CFE debuted in 1981 for translation of mRNA, and the platform has since been optimized for combined transcription and translation with batch yields ∼50 μg/mL68 and semicontinuous yields ∼1 g/L67 with the capacity to produce functional antibodies.107,108 Less productive platforms include HeLa,69 hybridoma,109 and historical systems like PC12110 and mouse111 or rat liver9 cell extracts. These CFE systems can facilitate high-throughput prototyping at small scales prior to application in cultures of therapeutic-producing cell lines.112
While other eukaryotic systems, such as those derived from fungi,74,113,114 trypanosomes,115 and insects75,76 are also being developed, the field is still working to develop strategies to reduce costs, increase yields, and efficiently produce proteins with humanized glycosylation patterns.
2.2. Cell Extract Preparation Methods
The preparation of cellular extracts is integral to efficient protein biosynthesis in vitro. The workflow for producing extracts generally consists of: (i) growing and harvesting cell cultures, (ii) lysing cells, and (iii) processing cell extracts. However, the broad applications for CFE and adoption by various research laboratories have resulted in subtle differences within each step of extract preparation. In this section, we detail methods to produce extracts and highlight certain steps which may be necessary for certain CFE applications. Throughout this section and those that follow, we focus on bacterial CFE systems derived from E. coli.
2.2.1. Extract Preparation from E. coli
Cell extracts can be prepared using a variety of E. coli strains and processing steps to enhance performance in a desired application (Figure 2). State-of-the-art CFE workflows commonly use engineered derivatives of the BL21 strain, which are typically high-yielding due to enhanced mRNA stability and decreased protease activity. Specific applications may call for other strains that are described in relevant sections and/or reviewed previously.20 For example, using engineered E. coli strains optimized for disulfide bond formation within the cytoplasm is useful for synthesizing full-length antibodies within CFE.116 In addition, E. coli extracts can be made from derivatives of the genomically recoded C.321ΔA strain to incorporate noncanonical amino acids (ncAAs) into proteins more efficiently.117
Figure 2.

Cell extract preparation workflow. All protocols incorporate growth and lysis of cells, with variation in the density and lysis method. The soluble fraction of lysate is referred to as extract, and this may be further processed with a runoff incubation, dialysis, or lyophilization for strain- or application-dependent increases in activity.
Once the strain is selected, it is then grown and harvested. Typically, freshly isolated colonies from agar plates are used, although glycerol stocks are an alternative.118,119 Cells are generally grown using common media formulations like Luria–Bertani (LB) broth, 2x Yeast Extract Tryptone (2x YT), 2x Yeast Extract Tryptone Phosphate (2x YTP), and 2x Yeast Extract Tryptone Phosphate Glucose (2x YTPG). Protein synthesis yields could be improved by using 2x YTPG media relative to LB media from 37% to 500% in conjunction with CFE conditions that mimicked the E. coli cytoplasm.120,121 In addition, the inclusion of glucose in the growth media is hypothesized to upregulate genes involved in central metabolism, leading to improved protein synthesis capabilities while lowering phosphatase activity.19,121 Glucose, however, was found to inhibit translation from endogenous bacterial promoters.122 This suggests that the choice of media should be dictated by the final application for CFE.
Protein expression during cell growth can enrich heterologous proteins of interest within the cell extract. These enriched extracts are useful when a protein of interest is necessary for a downstream application or when it simplifies the reaction setup. Both constitutive and inducible expression systems can be used, but inducible systems are often preferred, particularly when the protein of interest is deleterious to cell growth. In one example, the T7 RNA polymerase-promoter system is used and is induced from the lac operon in BL21 (DE3)-based strains to circumvent the need to separately purify and supplement T7 RNA polymerase into CFE reactions. Methods using autoinducing media compositions that enable protein expression without user intervention during cell extract preparation have also been developed.87,123,124 This reduces the labor required for extract preparation and improves the total yield of cell extract from a single preparation.
Cells are typically harvested by centrifugation during mid-to-late exponential phase. This maximizes the yield of cell extract and translation machinery components recovered while maintaining productivity.125 Extracts made from cultures at high cell densities are less productive than using traditional extract preparation protocols.87 This decrease in productivity is hypothesized to result from a smaller number of active ribosomes.126 However, optimized media formulations with increased buffering capacity can prevent acidification of the cell culture and maintain extract productivity from high cell density cultures, resulting in four times the volume of extract without losing extract productivity.87 After harvesting, the cells can be processed into extract immediately or stored at −80 °C until a more convenient time.
Crude cell extracts are prepared by lysing cells, which to date has been achieved using homogenization,127−130 sonication,125,131 bead-beating,131,132 or enzymatic lysis.133,134 Cell lysis is the step most likely to differ between laboratories that have adopted CFE. Homogenization and sonication are most widely used for extract preparation and historically have produced the highest-yielding extracts. However, the choice of lysis method depends on variables such as the volume of cells, the available instrumentation, the presence of membrane vesicles,135 and the E. coli strain used. For example, processing large volumes of cells benefits from using homogenization to reduce the time invested during cell lysis.
After lysis, crude cell extracts are processed to remove cellular debris by centrifugation. Historically, extracts have been centrifuged at 30,000g, but slower centrifugation speeds (12,000g) have been found to improve extract productivity, which could be due to the increased number of inverted membrane vesicles that house enzymes involved in ATP regeneration.19,133,135 Lower centrifugation harvesting is also useful in applications requiring expression of membrane proteins.135 After centrifugation, extracts may be used immediately, aliquoted and flash frozen for later use, or further processed. Further processing can include dialysis, which removes small molecule inhibitors of transcription and translation, or a runoff reaction, which is hypothesized to provide ribosomes time to finish translating native transcripts as endogenous mRNA is degraded.136 The runoff step also reduces the abundance of several proteins that precipitate out of solution.137 Interestingly, these processing steps appear to be most important for transcription from native promoters and do not typically affect extract productivity when using T7 RNA polymerase in BL21 Star (DE3) strains.122,129 Hypotheses around these impacts include the depletion of cold shock-like proteins during runoff that otherwise hinder gene expression and/or the depletion of unknown inhibitory metabolites during dialysis.122,137 Although more extensive analysis is required to understand the complex changes in extract-based systems during runoff and dialysis, it is clear that these post-lysis processing steps can be important for reconstituting activity in specific applications of CFE.
2.2.2. Lyophilization of Cell Extract
Creating shelf stable CFE reactions using lyophilization is a growing area of research. This may have promising applications in on-demand production of therapeutics,138−141 biosensors142,143 at the point-of-use, and educational kits.144,145 Compared to fresh reactions that lose activity after storage at room temperature over a month, lyophilized CFE reactions retain protein synthesis activity in equivalent conditions over several months.140,146 Extract stability after lyophilization can be enhanced with additives. CFE reactions that are lyophilized with sugars like trehalose and lactose, for example, retained more protein synthesis activity compared to those lyophilized without additives after 2 weeks.147,148 Lyophilized CFE systems have been reported to maintain activity out to 90 days,140 and the longevity of lyophilized CFE reactions past that are still being actively explored. Additives are particularly appealing in combination with the observation that sugars such as maltodextrin can be used as alternative, cheaper energy sources in CFE.40,45
2.2.3. Cell Extract Preparation from Other Organisms
While the previous two subsections focused on E. coli extracts, extracts have also been prepared and characterized from over 40 different organisms. Extract preparation for these organisms differ from E. coli-based methods on a case-by-case basis. For example, active extracts from the fast-growing marine bacterium Vibrio natriegens has improved protein synthesis yields from cells grown in brain-heart infusion broth compared to the more common 2x YTPG medium.64,92,93 Additionally, compared to homogenization pressures used to lyse E. coli (∼20,000 psig), the homogenization pressure to make extracts from Streptomyces lividans B-12275 were optimal at 12,000 psig.59 For eukaryotes such as S. cerevisiae, parameters such as OD of the harvest, RNase treatment, and others differ significantly from E. coli methods.149 Other differences in extract preparation methods for non-E. coli organisms have been reviewed elsewhere.20
2.3. Reaction Components and Energy Systems
In this section, we describe the components most frequently found in cell-free reaction mixtures to activate metabolism, transcription, and translation (Figure 3). We focus on formulations developed for contemporary E. coli-based systems; a more detailed historical perspective has been reviewed elsewhere.33
Figure 3.

Typical additives to CFE systems. Components required for protein synthesis are grouped based on the three main processes activated in the reaction: metabolism, transcription, and translation. Components highlighted in blue represent factors that are supplemented while components highlighted in red are either present endogenously or can be supplemented into the cell-free reaction depending on the formulation.
2.3.1. Energy and ATP Production
Activating metabolism in a CFE reaction is necessary to enable regeneration of ATP to fuel high-level transcription and translation. CFE formulations either activate the native metabolism of the cell extract or rely on supplemented metabolic pathways. These metabolic pathways are supported by energy inputs, cofactors, buffers, and salts in numerous formulations.
Phosphorylated energy substrates with high-energy phosphate bonds such as phosphoenolpyruvate (PEP),150,151 creatine phosphate,26,152,153 or acetyl phosphate154 were traditionally used as CFE substrates in combination with their respective kinases to generate ATP. Unfortunately, these expensive phosphorylated substrates typically account for the majority of reaction cost, which has driven innovation toward more economical alternatives.155−157 Additionally, phosphorylated substrates are degraded by phosphorylases present in the extract, destabilize reaction pH,18 and cause the accumulation of inorganic phosphate that can sequester magnesium ions necessary for protein synthesis.158,159
To alleviate these limitations, other energy substrates have been investigated in E. coli-based systems. For example, ATP generation from pyruvate can be activated without additional enzymes by supplementing reactions with nicotinamide adenine dinucleotide (NAD) and coenzyme A (CoA).151 Other glycolytic intermediates like glucose-6-phosphate (G6P),151 fructose-1,6-bisphosphate (FBP),155 and 3-phosphoglycerate (3-PGA)42,160 can also be used as energy sources for CFE. Low-cost primary energy sources such as glucose,40,156 maltose,161 maltodextrin,148,161,162 and soluble starch41 can be fed into glycolysis, but require the addition of a phosphate source (e.g., potassium phosphate40,41 or hexametaphosphate (HMP)161).
The discovery that early glycolytic intermediates could be used to fuel CFE also resulted in new systems combining multiple energy substrates and metabolic pathways in the extract. For example, the PANOx system includes NAD and CoA in addition to PEP, which increases reaction productivity by harnessing both ATP from the dephosphorylation of PEP as well as the downstream metabolism of the resulting pyruvate.151 In another example, the 3-PGA system has been coupled with both maltose and maltodextrin to manage the accumulation of inorganic phosphate, which was found to extend the reaction time and resulted in high yields.163
In addition to glycolysis, the tricarboxylic acid (TCA) cycle and oxidative phosphorylation can be activated in the cell-free context. In an effort to optimize cell-free reaction productivity by more closely mimicking the cellular environment, the PANOx-SP system removed unnatural components such as buffer and polyethylene glycol (PEG), adjusted salt concentration, and added in the polyamines spermidine and putrescine.18 While this did not make a large difference in protein synthesis levels when PEP was used as an energy substrate, it dramatically improved the yields and maintained consistent pH when pyruvate was added (Cytomim system) as the energy substrate instead of PEP.18 This system was also active in the absence of pyruvate, indicating that the Cytomim system supported other modes of energy generation.120 Glutamate was also described to be an energy source, which can produce NADH through the TCA cycle and fuel oxidative phosphorylation enabled by the presence of a proton gradient and ATP synthase found in inverted membrane vesicles in the extract.19 Doubling the glutamate concentration and optimizing other reagents can also result in a robust and low-cost protein synthesis system.157
2.3.2. Cofactors, Small Molecules, Ionic Strength, and pH
The cofactors NAD and CoA are expensive reaction components that are desirable targets for removal to decrease reaction cost. Supplementation of NAD was crucial for the metabolism of pyruvate in a traditional S30 extract formulation, and CoA further boosted protein synthesis yields.151 Removal of NAD and CoA resulted in substantial decreases in protein synthesis yields in subsequent optimizations using glucose as an energy source.156 However, changes to expedite extract production such as low-speed clarification steps (S12) and the absence of the dialysis step, have changed the concentration of endogenous components in the final extract.125,130 Optimizations using glutamate,157 FBP,155 and maltodextrin148 have found supplementation of NAD, CoA, or both cofactors to be unnecessary in the reaction when dialysis is not carried out.
Maintaining the overall ionic strength of the CFE environment is important to mimic the cellular environment and maintain the proper function of biological processes. Cations such as magnesium, potassium, and ammonium are traditionally added to CFE reactions to balance the charge of negative biomolecules, like nucleic acids.33 Magnesium plays a significant role in CFE productivity as it is involved in ternary complex formation, ribosome assembly, and tRNA aminoacylation, among others, so the magnesium concentration is typically optimized for each individual extract.14 Importantly, due to the high necessary concentration of supplemented cations, the accompanying anionic counterions, often glutamate or acetate, are present at higher concentrations in a CFE reaction than in the E. coli cytoplasm.18 Formulation iterations use glutamate rather than acetate or chloride salts to better reflect the cytoplasmic environment and reduce acetate concentration.18
Small molecules are also added to improve energy regeneration. Oxalic acid is a commonly added small molecule that inhibits PEP synthetase and avoids the unproductive conversion of pyruvate back to PEP.164 However, oxalic acid is detrimental to protein yields when using early glycolytic intermediates such as G6P as an energy source.151 Phosphate salts such as potassium phosphate are also important when using early glycolytic substrates such as glucose or maltodextrin for initial substrate phosphorylation.40,41
Maintaining appropriate pH in the reaction is important for metabolism and many other CFE processes. While the Cytomim system removed the HEPES buffer entirely,18 certain energy sources, like glucose, have different impacts on reaction pH, typically due to the production of acidic byproducts.40 Important considerations for buffer use in CFE include buffer concentration, pH, and pKa/buffering range. HEPES at pH 7.2–8.2 is most commonly used; however, it was found that Bis-Tris (which has a lower pKa than HEPES) improved protein synthesis when using glucose-based energy systems and better stabilized reaction pH.40
2.3.3. Transcription
RNA polymerase (RNAP) is required to transcribe the DNA encoding the gene of interest into RNA in CFE reactions. Bacteriophage polymerases such as T7 or Sp6 polymerases are often used16,99 and can be supplemented as a purified enzyme or induced during cell growth and enriched in the extract.128 The endogenous E. coli RNA polymerase and sigma factor 70 (σ70) can also be leveraged for transcription in CFE, although optimal reaction formulation and extract processing conditions required for efficient expression can differ from bacteriophage RNAP systems.122,165
For systems that incorporate more complex transcriptional programs, the addition of exogenous regulatory elements is required. For example, IPTG42 and cyclic adenosine monophosphate (cAMP)33 can be supplemented to the cell-free environment to induce regulated gene expression. Transcription factors are also commonly added or enriched in the cell extract to build genetic circuits or biosensors, which is discussed further in Sections 3.5 and 3.6.
Nucleotides are critical additives for transcription. Nucleoside triphosphates (NTPs) are the most common form of nucleotide supplement; however, they are one of the most expensive additives. It has been demonstrated that the native cellular machinery in extract can phosphorylate nucleoside monophosphates (NMPs) when they are added to the reaction instead of NTPs, which is more cost-effective.19,156,157 In fact, protein synthesis can be carried out without the supplementation of exogenous nucleotides in some formulations.148,166
Crowding agents can be added to CFE reactions to help stabilize molecules such as RNA and promote macromolecular interactions. These additives can impact both transcription and translation, and their identity and concentrations are an important factor to consider as the protein concentration in a CFE reaction is approximately 20 times lower than the protein concentration in the cytoplasm of E. coli.18 PEG is commonly added as a non-native crowding agent, but optimal concentrations can depend on the molecular weight of PEG38,167,168 and the desired protein product.44 Other crowding and stabilizing agents that have been investigated in addition to PEG include alcohols, sugars, amino acids, and polyols.167 While PEG is still used in some contemporary formulations, the Cytomim formulation notably replaced the unnatural crowding agent with the native polyamines spermidine and putrescine to stabilize RNA and other compounds,18 which can also impact translation (see Section 2.3.4).
2.3.4. Translation
Transfer RNA (tRNA) recognizes the mRNA sequence and brings the corresponding amino acids to the ribosome to enable translation. Typically, E. coli tRNA is supplemented to the cell-free reaction; however, it is not an essential additive due to its native presence in the extract. Optimizations using FBP,155 glutamate,157 and maltodextrin45 as energy sources have removed exogenous tRNA without a negative impact on protein synthesis, suggesting that native tRNA is present at suitable concentrations in the S12 extracts used.
Folinic acid is a necessary precursor to N-formylmethionine and has been historically supplemented due to its importance in translation initiation. However, exogenous folinic acid was determined to be unnecessary for high levels of protein synthesis in a glutamate-based CFE system.157
Typically, the 20 amino acids are supplemented in excess to the CFE reaction at mM concentrations. Metabolism in the extract can both consume and produce amino acids. For example, depending on the extract source strain, it has been demonstrated that aspartate, lysine, and tyrosine can be produced in the cell-free reaction,19 so amino acid concentration could be a point of optimization or target for removal from the reaction in future studies.157 Investigation into the necessary amino acids for a specific protein of interest could result in reduced cost for a particular product. Optimized concentrations of serine and glutamine, for example, led to increased yields in a creatine phosphate fueled system.169
Dithiothreitol (DTT), a reducing agent used to stabilize proteins with free sulfhydryl groups, is a common additive during extract preparation when a reducing environment is desired during CFE.125 Optimal concentrations of DTT in CFE systems generally are around 1 mM.33,165,167
The polyamines spermidine and putrescine are added to the Cytomim formulation to more closely mimic the cytoplasmic environment.18 They also facilitate ribosome assembly.170 While high concentrations of putrescine are found in E. coli cells, optimizations found that when adding spermidine, putrescine is not needed in the cell-free reaction.157
2.3.5. Expression Templates
The DNA or RNA templates used to drive CFE take a variety of forms (Figure 4). The uniting feature of these templates is that they contain the appropriate elements to activate transcription (e.g., promoter, terminator, etc.) and/or translation (e.g., ribosome binding site (RBS), Kozak sequence, open reading frame (ORF), etc.) in the CFE system of interest. Typically, synthetic constructs are built containing the gene of interest and the required regulatory elements,171 although genomic DNA and RNA, for example phage genomes,172 are also used as templates for CFE. For protein expression applications, it is common to use vectors lacking inducible regulatory elements (e.g., the lacO site) in order to maximize protein expression yield.173
Figure 4.

Templates used to drive transcription and/or translation in CFE systems. IVT: in vitro transcription; DNA: deoxyribonucleic acid; mRNA: messenger ribonucleic acid; PCR: polymerase chain reaction; RCA: rolling circle amplification.
Users of CFE systems are faced with the choice of using either DNA or RNA as templates to drive gene expression. Here, in vitro transcription (IVT) derived mRNA is commonly used to drive translation-only CFE reactions and DNA is used to drive combined, one-pot transcription and translation reactions. Several considerations can help users choose between the two templates. Combined reactions (note transcription rates and translation rates are not typically coupled/matched and thus we use the word combined) tend to be more convenient, as they require fewer steps to prepare the template and do not require the manufacture, handling, and storage of RNA. Many common CFE systems can perform high yielding, combined transcription and translation. These include the prokaryotic E. coli,21,23 the protozoan L. tarentolae,70,174 the insect S. frugiperda,175 the plant N. tabacum BY-2,71 and the mammalian CHO107,176 CFE systems. However, in some common CFE systems (e.g., wheat germ,78 rabbit reticulocyte177), mRNA is the typical template, although there are combined versions available.178 Combined (i.e., transcription and translation) and translation-only reactions can also have different yields. In some systems, transcription and translation can have different optimum conditions (e.g., temperature, salt) or cause resource competition that reduces yields, which can be overcome with mRNA as a template.177 However, in high yielding E. coli systems, IVT mRNA has been found to result in lower protein yields than its DNA counterpart.179,180 This was hypothesized to be related to mRNA folding leading to inaccessibility of the translation start site,179,180 and interestingly this issue was partially or completely alleviated by performing the IVT step in the presence of a ribosomal extract.180
Linear DNA templates, frequently referred to as linear expression templates (LETs) (reviewed elsewhere181−183), are particularly attractive because they circumvent the time-consuming and low-throughput steps associated with plasmid assembly, cell culture, and plasmid purification, saving days of effort and enabling throughputs much greater than cell-based workflows.184−187 High-throughput LET approaches often involve an assembly step, to build an LET capable of supporting transcription and translation, and an amplification step, to manufacture the template in sufficient quantity to drive a CFE reaction. The assembly step typically involves assembly of the gene with other regulatory elements (e.g., promoter, ribosome binding site, terminator, etc.) and potentially with other components of the ORF (e.g., different protein fragments). To date, assembly has been accomplished with Gibson assembly,185,186,188,189 Golden Gate assembly,185 Gateway assembly,187 or polymerase chain reaction (PCR) based assembly.32,34,190−192 After assembly, LETs are amplified to generate sufficient quantities to drive a CFE reaction either by PCR32,185,187,191,193−195 or rolling circle amplification (RCA).184,187,196,197 PCR generates linear dsDNA containing only a single copy of the template and requires thermal cycling, whereas RCA leverages a circular template, exonuclease resistant primers, and a strand displacing polymerase to generate large fragments of mixed ssDNA and dsDNA containing many copies of the template in a single molecule. Because the assembly and amplification reactions are in vitro enzymatic reactions, they can be carried out sequentially in microplates and without cells, making them amenable to high-throughput screening applications.
A major consideration in the use of linear DNA in CFE is that crude extract-based systems contain exonucleases that degrade LETs, which can hinder protein expression. Interestingly, cell culture conditions and extract processing steps play a major role in the ability of a given CFE system to utilize LETs. For example, when using E. coli extracts that are not processed with runoff or dialysis, protein yields using LETs can be within 50% of that from plasmids.116 Methods that protect LETs from degradation are frequently required. The RecBCD exonuclease complex is known as a contributor to the instability of linear DNA in E. coli(198) and has been the major target for inhibition in E. coli CFE systems. Strategies tested to inhibit RecBCD include genomic modification,193,199,200 competitive substrate inhibitors,201 protective DNA sequences,202 protein inhibitors,185 and small molecule inhibitors203 which have yielded varying degrees of success and are summarized in Table 4. The success inhibiting the RecBCD exonuclease in E. coli has led to other strategies to inhibit exonucleases more broadly that may be applicable in other CFE systems. Several different DNA modifications with potential to increase nuclease resistance, including methylation204 and chemical modification,185,200 have been evaluated but yielded little to no improved protein expression. A more successful general strategy has been to utilize DNA binding proteins that block the terminal ends of the LET,204−206 which in one case resulted in LET-based expression on par with that of plasmid-based expression.205 Circularization of the linear template DNA has also been a successful strategy for protecting against degradation, yielding similar results between linear and plasmid-based templates.207 A potential benefit of RCA-derived templates over PCR-derived templates is that they are reported to be less sensitive to degradation, likely a result of the long length of the LET product serving as its own competitive substrate inhibitor.196
Table 4. Strategies for Protecting Linear DNA in CFE.
| Strategy | Implementation | Proposed Mechanism | Result | Organism | Reference |
|---|---|---|---|---|---|
| Extract preparation method | No runoff or dialysis | Unknown | 50% of plasmid yielda | E. coli | Hunt et al.116 |
| Genomic modification | A19ΔrecCBD::Plac-red-kan-ΔendA | Replacement of RecBCD with bacteriophage λ red recombination system | 45–63% of plasmid yield | E. coli | Michel-Reydellet et al.193 |
| Genomic modification | SBP tagged RecD and PNPase | removal of RecD and PNPase from extract | 2–4x improved LET yield | E. coli | Seki et al.199 |
| Genomic modification | ΔrecCBD | Knockout of RecBCD | 48–78% of plasmid yield | E. coli | Batista et al.200 |
| Nuclease inhibition | Chi6 oligos | Competitive inhibition of RecBCD | 23% of plasmid yield | E. coli | Marshall et al.201 |
| Nuclease inhibition | gamS | Protein inhibitor of RecBCD | 38% of plasmid yield | E. coli | Sun et al.185 |
| Nuclease inhibition | CID 697851, CID 1517823 | Small molecule inhibitor of RecBCD | 2.5–3x improved LET yield | E. coli | Shrestha et al.203 |
| Nuclease inhibition | Tus-Ter | Protein binds linear dsDNA termini | 100–146% of plasmid yield | E. coli, V. natriegens | Norouzi et al.205 |
| Nuclease inhibition | Ku | Protein binds linear dsDNA termini | improvement in transcription yield | E. coli and 4 nonmodel organisms | Yim et al.206 |
| Nuclease inhibition | ssCro | Protein binds linear dsDNA termini | 2–28% of plasmid yield | E. coli, V. natriegens | Zhu et al.204 |
| DNA modification | Terminal 5′ phosphorothioate linkages (x2) | 5′ DNA more resistant to nuclease degradation | 36% improved LET yield | E. coli | Sun et al.185 |
| DNA modification | Methylation | Methylation of linear DNA potentially inhibits degradation | No significant improvement | E. coli | Zhu et al.204 |
| Template design | Flanking DNA | 5–500 bp of flanking DNA as a buffer for exonucleases | 2.4–6x improved LET yield | E. coli | Sun et al.185 |
| Template design | Nuclease Resistant Sequence | Reduced activity of nuclease on linear DNA | 75% of plasmid w/o GamS and 100% of plasmid w/GamS | E. coli | Chen et al.202 |
| Template design | Circularization | PCR-based circularization of LET | same as plasmid yield | E. coli | Wu et al.207 |
| Template amplification method | Rolling Circle Amplification | Long concatemers serve as competitive inhibitor for exonucleases | 75–80% of plasmid yielda | E. coli | Gyanendra et al.197 |
Approximate result listed when the exact improvement was not reported.
There are several other factors to consider when choosing a template for CFE. Template “quality”—both the purity of the template and the damage accumulated by the template during purification—are known to impact CFE reactions and lead to increased variability, and thus preparation methods can have an impact on CFE results.208,209 Toward improving template preparation outcomes, best practices for template preparation have been suggested.210 Fortuitously, in applications utilizing linear DNA, the buffers associated with both PCR and RCA have been found in some cases to be tolerated in CFE, enabling the use of unpurified template.116,187,188,196 Methods for quantifying the concentration of unpurified DNA templates have also been reported.196
Linear DNA may not be suitable for all applications. In efforts to evaluate E. coli σ70 promoters in CFE, multiple researchers have observed that the strength of a given promoter correlated better with in vivo results for plasmid-based CFE than for LET-based CFE.185,211 It was hypothesized that this was related to DNA supercoiling and its impact on transcription rate.211 Interestingly, others have found that the addition of linearized vector instead of PCR product eliminated the difference between linear and circular templates when evaluating the transcription factor TetR.212 The authors proposed that nonspecific binding of transcription factors to the vector could be the a cause of the observed differences.212 Thus, it is important to evaluate the suitability of linear templates for a given application before undertaking large scale experiments. While this may be a limitation for some applications for the time being, advances in cell-free plasmid replication systems may enable enzymatic plasmid amplification with the same ease as the amplification of linear DNA.213,214
2.4. Purified CFE Systems
While the previous subsections have covered extract-based CFE systems, in vitro transcription and/or translation can also be recapitulated using only purified components (Figure 5), allowing the reaction composition to be precisely defined. The state-of-the-art and widely used purified CFE system is called the PURE system.26,215−217 PURE comprises purified transcription and translation machinery mixed with a circular or linear DNA template and energy sources to activate transcription and translation. The proteinaceous components of PURE are sourced from E. coli and consist of all 20 aminoacyl-tRNA synthetases (aaRS), Initiation Factors (IF) 1–3, Elongation Factors (EF) Tu, Ts, and G, Release Factors (RF) 1–3, Ribosome Recycling Factor, Methionyl-tRNA formyltransferase (MTF), T7 RNA polymerase, creatine kinase, myokinase, nucleoside-diphosphate kinase, and pyrophosphatase. The system is further supplemented with E. coli tRNAs and ribosomes. Finally, PURE also includes relevant small molecules essential for transcription, translation, and energy regeneration, including all 20 amino acids, 10-formyl-5,6,7,8-tetrahydrofolic acid, NTPs, and creatine phosphate. PURE is commercially available as PURExpress (NEB) and PUREfrex (Gene Frontiers), which has enabled its use by many laboratories worldwide.
Figure 5.

CFE systems using purified components. (A) Key components required for purified CFE systems. (B) Methods of obtaining purified CFE systems.
The differences between CFE using PURE or cell extracts mainly arise from the presence of proteins and small molecules within the E. coli cytoplasm that remain after extract preparation. Functionally, this has resulted in reduced capacity in PURE reactions, relative to extracts, in terms of number of proteins produced per ribosome.218 This must be considered as these components can affect its performance in a desired application. We have listed several general differences in Table 5.
Table 5. Comparison of PURE and Extract-Based CFE Systems.
| Category | Extract-Based | PURE | |
|---|---|---|---|
| Reaction Composition | Redox State | Typically reducing. Can be manipulated by addition of small molecules such as glutathione. Modified redox state may benefit from genomically modified strains. | Typically reducing. Can be manipulated by addition of small molecules such as glutathione. |
| Energy Regeneration | Many energy sources, including glucose, PEP, and others, can be used due to active central metabolism. | Uses creatine phosphate and creatine kinase system to regenerate ATP and cofactors. | |
| DNA Templates | Both plasmid and linear DNA can be used. Linear DNA may require additional protection methods depending on the CFE system. | Both plasmid and linear DNA can be used. | |
| Vesicles | Already present in cell extracts.135 | Must be supplemented if necessary for specific application.219 | |
| RNA Polymerase | Both T7 and E. coli RNA polymerase are commonly used. | T7 RNA polymerase most common. E. coli RNA polymerase is less active and requires additional factors.220 | |
| Practical Considerations | Maximum Batch Yield | >2–4 mg/mL protein.21,23 | ∼0.1 mg/mL to 1 mg/mL protein.221,222 |
| Difficulty of Making Reagents | Easy to Moderate. Extracts optionally require processing after lysis. | Difficult. Proteins and ribosomes must be purified from cell extracts. | |
| Commercial Availability | Widely available (e.g., myTXTL (Arbor Biosciences) and NEBxpress (NEB), among others). | Widely available (PURExpress (NEB) and PUREfrex (Gene Frontiers)). | |
| Cost | ∼ 5 $/mL homemade1 and ∼400 to 1,000 $/mL commercial.223,224 | ∼100 $/mL homemade225,226 and ∼ $1,000 to 1,200 $/mL commercial.221,222 | |
| Applications | Short Peptide Synthesis | Peptides can be unstable in extract-based systems due to low translational efficiency or proteolytic degradation227 unless attached to a fusion protein.228,229 | Peptides are stable in purified systems. |
| Modeling | Models must be fit empirically due to unknown composition of extract components. | Well-defined composition enables bottom-up modeling | |
| Noncanonical Amino Acid Incorporation | Amber suppression using orthogonal translation systems is highly productive22,23 | Genetic code reprogramming using flexizymes and traditional orthogonal translation systems230 | |
Historically, PURE has been made by individually purifying all transcription and translation machinery from E. coli overexpression cultures and mixing them at defined concentrations.26 While this affords the greatest amount of control over the reaction composition, this process is time-consuming, laborious, and expensive. There have been several efforts to simplify the preparation of PURE. Most examples focus on reducing the number of purifications required by including multiple PURE components within a single culture. This has been accomplished by installing genomic His tags onto PURE proteins within a set of E. coli strains,231 grouping PURE components onto multicistronic plasmids,232 or by coculturing all 36 E. coli strains required to express each PURE component within a single culture.225,226 One challenge of these systems has been batch to batch variability, which was recently addressed by modifying promoters for more stable expression.233 While all methods that have moved away from individual protein purification sacrifice some control over PURE’s composition, they circumvent the high cost and laborious nature of producing PURE, which are two of the main barriers to its use. With further improvements, the preparation of PURE could perhaps become as simple as the preparation of extracts for in vitro transcription and translation.
As protein yields from PURE are lower than those from extract-based CFE systems, there have been many efforts to optimize protein yields from PURE. Combinatorial optimization of transcription and translation components has shown that increasing the concentration of translation factors increases protein synthesis capacities.234,235 Changes to improve PURE can also be guided by computational models, which are more tractable for PURE compared to extract-based systems.236 Another direction has focused on the addition of supplements to PURE. Similar to extract-based CFE, the addition of crowding agents in PURE to mimic the cellular environment has also improved protein yields.234,237 A recent example has shown that incorporating dextran allows reduction of the proteinaceous components of PURE by up to 97.3% while maintaining protein synthesis yields, allowing PURE to be used more effectively.238 To prevent ribosome stalling, a known issue in PURE, elongation factors, such as EF-P, GreA, and GreB, can be added to improve protein synthesis.220 It has also been reported that the addition of ribosomal protein S1 improves PURE by increasing the number of functional ribosomes.239 The addition of protein chaperones, such as GroEL/ES, can also be useful for proteins that are difficult to fold.234,240 Additionally, fed-batch reactions where depleted small molecule substrates are added back into the reaction can also improve yields from PURE. Finally, PURE can be improved by using different energy regeneration systems. For example, the use of a polyphosphate kinase to regenerate ATP and GTP in place of the three kinases present in conventional PURE increases yields of active protein from ∼400 μg/mL to >500 μg/mL of sfGFP.241 These improvements are useful for making PURE more cost-effective and efficient for protein synthesis.
2.5. CFE Reaction Formats and Scales
CFE reactions can be grouped by the function(s), reaction vessel, and format for supplying reagents (Figure 6). Reactions may be transcription-only, translation-only, or combined transcription and translation, offering varying levels of control.1 Reactions are often executed in batch mode, where the reagents described above are combined and incubated until a limiting reagent is exhausted or inhibiting products accumulate. Batch reactions are versatile and can be performed across multiple scales including: encapsulated emulsions and vesicles at the femtoliter to nanoliter scale, tubes or well-plates at the microliter scale, thin films and test tubes at the mL scale, and bioreactors at the mL to L scale.29,37,242 However, batch reactions have limited longevity and output due to the accumulation of metabolic byproducts and exhaustion of energy sources. These limitations are addressed in fed-batch and continuous exchange (also referred to as semicontinuous) setups in which additional template, substrates, and/or energy sources are provided during the reaction,29 increasing reaction longevity and/or yield.21,153 Continuous reactions (either continuous exchange175 or continuous flow243) are often carried out in dialysis membranes74 or bilayers42 such that the CFE reaction is surrounded by a feeding solution to increase the longevity and productivity of gene expression.28 Alternatively, semipermeable liposomes and microfluidic chips can function like dialysis cassettes to enable continuous reactions.21,244,245 In all formats, aeration is a significant consideration for sufficient activity of oxidative phosphorylation to generate ATP for transcription and translation.19
Figure 6.
CFE reaction formats and scales. Cell-free reactions carry out metabolism, transcription, and/or translation as a part of gene expression. CFE is perfromed in batch or continuous modes in a range of vessel sizes and geometries.
Additionally, CFE reactions are amenable to incorporation within material matrices, which offer unique options for point-of-use applications and functionalized biomaterials. Key examples include clay mineral gels for studies mimicking early life,246,247 fibrous matrices like paper for distributable biosensors,248−250 diverse hydrogels for reaction immobilization,251−254 and polymer encapsulation for printing sturdy structures imbued with biological functions.255−257 Two recent reviews focus more extensively on the combination of CFE with materials.258,259
2.6. Monitoring and Quantifying the Results of CFE
CFE systems are complex mixtures, which can complicate quantification of the products of their reactions. Evaluating the functional activity of the expressed molecule (e.g., fluorescence, binding, catalytic activity, etc.) or system of molecules (e.g., complex assembly, metabolic pathway productivity, etc.) is frequently the goal; however, quantifying the results of transcription and translation in the system is often crucial to understanding the functional result. Here, we review methods for quantifying RNA and protein produced by CFE systems.
Incorporation of nucleotide radiolabels into RNA74 and amino acid radiolabels into proteins260 is the gold standard to quantify the molecules produced during the CFE reaction and does not require modification of the expressed sequences. To quantify transcription, it is common to fuse a fluorescent aptamer to the RNA of interest.22,51,59,261,262 Similarly, protein fusions — including fluorescent proteins,187,263,264 high affinity luciferase complementation reporters,95,265 tetracysteine tags,266−268 and purification or antibody epitope tags for Western blotting95,135,141,269,270 — have been used to quantify the products of translation. Other fusions have been suggested,271 including a fluorescent protein complementation reporter,272 protease-cleavable fluorescent proteins,187,264,273 and nanoluciferase.274
Several methods to track RNA or protein production in CFE systems have been developed that do not require modification of the expressed molecule. Transcription in the PURE CFE system has been monitored without modifying the transcript using Förster resonance energy transfer (FRET) based binary DNA primer probes.275 As an alternative to radiolabels, tRNAs precharged with a noncanonical amino acid containing a trackable modification (e.g., a fluorophore or biotin) are also widely used to stochastically label the native protein sequences produced in CFE.276 Furthermore, standard protein visualization techniques like gel electrophoresis and staining can be used to visualize protein expression from an unpurified CFE reaction provided enough product is expressed.277
2.7. High-Throughput Experimentation for CFE Applications
CFE systems are valuable tools for accelerating biological research. Below, we highlight technologies that are often coupled to CFE systems to enable high-throughput experimentation.
The simplicity of setting up CFE reactions has led multiple groups to develop automated systems for the manufacture and purification of proteins. Early embodiments combined template and CFE reagents with automated systems to synthesize up to milligram quantities of purified protein.77,278−280 Systems have also been developed that leverage onboard DNA synthesis and amplification combined with CFE to manufacture purified protein directly from a digital DNA sequence within an integrated instrument.32,191 These automated protein production systems point to a possible future where protein production is an entirely automated process with the only input from the user being the desired protein sequence, analogous to the advancements made in the DNA synthesis industry.281 Even when implemented without automation, the process of producing 0.1–1 mg of purified protein via CFE, from generating DNA template to purifying the protein of interest, can require similar effort to completing a plasmid purification kit.282 These methods can save researchers time and enable them to evaluate more hypotheses given a fixed amount of effort.
Liquid handling robotics can be integrated with CFE systems and used for the rapid setup of hundreds to thousands of distinct conditions using microplates. Traditional tip-based liquid handling robots have been interfaced with CFE systems largely to systematically test hundreds of distinct reaction conditions,283−285 with recent iterations leveraging fully automated cloud laboratories.286 Acoustic liquid handlers (e.g., the Echo instruments developed by LabCyte and now sold by Beckman Coulter) have enabled flexible 384 well reaction setup in minutes with reaction volumes as low as 0.5 μL.287 Acoustic liquid handlers have thus far been used to optimize reaction conditions,288 improve experimental reproducibility,208,289 generate data for model parametrization,51,288,290 and enable high-throughput genetic part51,112,212 and protein95,186 prototyping. Custom microwell systems also highlight the potential for smaller reaction volumes and higher density microwells.291 While lower throughput than the other technologies discussed in this section, microplate and liquid handler-based screening has the significant advantage of requiring less specialized knowledge to implement and being highly adaptable to the application of interest.
Spatial segregation or compartmentalization on a surface is another common strategy for extending the capabilities of CFE systems. Several immobilization strategies using self-assembling microarrays have been implemented including the protein in situ array (PISA),292,293 the multiple spotting technique (MST),294 the nucleic acid programmable protein array (NAPPA),295−297 the DNA array to protein array (DAPA),298 and microintaglio printing (μIP).299,300 These strategies generally rely on spatial segregation of the template DNAs in a DNA array and noncovalent capture of the synthesized protein to the chip to manufacture the protein array. This allows proteins to be produced in situ immediately prior to experimentation and without having to express and purify each individual protein, two major benefits over traditional protein microarrays. In addition to the standard fluorescence-based readouts,301,302 self-assembling microarrays have also been analyzed using fluorescent single-walled carbon nanotubes303 and surface plasmon resonance (SPR) sensors.304
Analogous to protein microarrays, several microfluidic surface-associated compartment systems containing tens to thousands of distinct compartments for CFE have been developed. These systems have been used to measure gene expression and genetic circuits, molecular interactions,303,305−307 the self-assembly of proteins308 and ribosomal subunits,309 and enzymatic activities.310 A key benefit of this approach is that the information in each chamber is linked to a known genotype and can sometimes be monitored in real-time. Furthermore, fabricated chip devices offer exquisite chemical, spatial, and temporal control over the process of gene expression and protein assembly, which has enabled the detailed investigations of a myriad of genetic circuits,245,311−314 molecular machine assemblies,309,315−317 and synthetic cell systems.245,313,314 Some iterations of this technology enable millions of defined compartments,318 although the genotype is not known until individual wells are recovered. Recent work using cell-free displayed proteins spatially segregated on Illumina sequencer flow cells has enabled the collection of large (104–106 variants) data sets of quantitative and genotype-phenotype coupled measurements of binding proteins.319,320
Nonsurface associated compartments are also frequently used to encapsulate CFE systems. The methods vary widely in implementation and include emulsion-based droplets, liposomes, polymersomes, and more.29,321 These techniques can generate millions322−325 of compartments via vortex mixing, microfluidics, or other methods depending on the compartment type. Compartments are exciting prospects for CFE-based high-throughput screening and directed evolution,322−328 though unlike the surface-based microfluidic chambers above, they typically do not provide coupled information about the identity and activity of each condition screened. However, this is not the case for all applications. Droplet-based compartments combined with optical barcoding, the incorporation of unique combinations and concentrations of fluorophore barcodes, enabled the collection of millions of data points on phenotype-genotype coupled droplets to optimize genetic circuits.329 Compartments have also been used to study the impacts of compartmentalization on gene expression29,330−333 and to build toward synthetic cells29,334,335 (Section 3.9.3).
Several display technologies that utilize CFE have been developed.336 While cell-based display methods typically create the genotype to phenotype linkage through compartmentalization, CFE-based selection methods, like ribosome337 and mRNA338 display, frequently leverage a direct physical link between a displayed protein and its transcript to maintain the genotype and phenotype linkage. These methods circumvent the traditional transformation limitation with cell-based techniques and in principle enable library sizes of up to 1014 variants.339,340 Generally, these systems have been applied to the selection of proteins that bind a desired target (Section 3.4.2), although they have also been expanded to map protein–protein interaction networks,341−344 evaluate protein stability,345−349 and evolve certain classes of enzymes.350
2.8. Modeling CFE Systems
While CFE serves as an excellent platform for both product synthesis and prototyping for in vivo systems, extract-based CFE systems are complex mixtures with not all parts defined. Purified CFE systems, on the other hand, are well-defined but still involve many different components forming a complex biological reaction network.351 Models of CFE systems seek to better understand the activities and formation or degradation rates of essential components in CFE systems such as metabolic resources, inhibitory products, ribosomes, RNA polymerase, mRNA, and proteins. Due to the open reaction environment, CFE components can be tuned by varying the amount of DNA or other agents added to the system. Therefore, experiment-driven models can be constructed and validated to capture specific interactions such as production bottlenecks and predicting protein synthesis, which is reviewed elsewhere.352
Typical models are based on ordinary differential equations, using Michaelis–Menten and Hill-Langmuir equations to describe different components of system behavior. The granularity of the model determines how many equations are needed, with coarse grained models of protein transcription, translation, and degradation having as few as four equations353,354 and complex models of reaction networks351,355 requiring hundreds of equations. Models are often fit with experimental data to determine parameters, such as the parameters describing transcription and translation kinetics. While coarse-grained models are adept at broadly capturing trends in relative rates,353 with translation identified as a key target for overcoming experimental bottlenecks,290 discrepancies between published model-derived parameters and experimental results could be due to differences in extract-based systems between laboratories. The recent development of a toolbox for modeling CFE-based genetic circuits will lower the barrier to entry for future researchers.356
3. Applications of Cell-Free Gene Expression
CFE systems have been transformed by improvements in productivity, cost, scale, and complexity. Below, we highlight example application areas that take advantage of these improvements.
3.1. Expression of Proteins Using CFE
CFE systems make an ideal tool for the customized expression of proteins. To date, CFE systems have enabled the expression of disulfide bonded proteins, proteins that require folding chaperones, post-translationally modified proteins, membrane proteins, multi-subunit proteins, proteins containing oxygen sensitive cofactors, and more (Table 6). Here, we detail the efforts to use CFE to produce different types of functionally active, complex proteins.
Table 6. Examples of Cell-Free Synthesized and Functionally Active, Complex Proteinsa.
| Year | Protein | Protein Class | CFPS System | Yield (μg/mL) | Complex Elements | Reference |
|---|---|---|---|---|---|---|
| 1997 | α-Hemagglutinin scFv | Antibody | E. coli | 8.3 | folding chaperone required, disulfide bonded | Ryabova et al.357 |
| 2003 | α-Salmonella O-antigen scFv | Antibody | Wheat germ | 13 | folding chaperone required, disulfide bonded | Kawasaki et al.358 |
| 2004 | Urokinase protease | Protease | E. coli | 40 | disulfide bonded | Kim et al.359 |
| 2004 | vtPA | Protease | E. coli | 60 | folding chaperone required, disulfide bonded | Yin et al.360 |
| 2005 | scFv-GM-CSF fusion protein | Antibody Cytokine Fusion | E. coli | 43 | folding chaperone required, disulfide bonded | Yang et al.361 |
| 2005 | β2AR | GPCR | E. coli | ∼1,000* | membrane bound | Ishihara et al.362 |
| 2006 | IGF-I | Hormone | E. coli | 400 | disulfide bonded | Swartz363 |
| 2008 | HydA1 | [FeFe] Hydrogenase | E. coli | 22 | cofactors, oxygen sensitive | Boyer et al.364 |
| 2011 | rhGM-CSF | Cytokine | E. coli | 700 | folding chaperone required, disulfide bonded | Zawada et al.37 |
| 2011 | cIFN-α | Cytokine | E. coli | 400 | solubility issues | El-Baky et al.365 |
| 2011 | ATP Synthase | Molecular Machine | E. coli | NA | folding chaperone required, membrane bound, multisubunit | Matthies et al.366 |
| 2011 | HBc | Virus-Like Particle | E. coli | 125 | disulfide bonded, macromolecular assembly | Bundy et al.367 |
| 2014 | VH3–7/Vk3–20 IgG | Antibody | E. coli | ∼1,500 | folding chaperone required, disulfide bonded, multisubunit | Groff et al.368 |
| 2014 | ERBB2 (HER2) | Receptor Kinase | E. coli | 2 | membrane bound | He et al.369 |
| 2017 | α-NIP IgG | Antibody | CHO | 114* | folding chaperone required, disulfide bonded, multisubunit, glycosylated | Martin et al.107 |
| 2017 | α-SMAD2 IgG | Antibody | CHO | 9* | folding chaperone required, disulfide bonded, multisubunit, glycosylated | Stech et al.108 |
| 2017 | HuNoV VP1-GII.3 | Virus-Like Particle | E. coli | 620 | macromolecular assembly | Sheng et al.370 |
| 2019 | α-HER2 IgG | Antibody | PUREfrex | 124 | folding chaperone required, disulfide bonded, multisubunit | Murakami et al.371 |
| 2021 | PD glycosylated with F. tularensis O antigen | Glycoprotein | E. coli | 20 | glycosylated | Stark et al.141 |
| 2021 | PD glycosylated with C. jejuni glycan | Glycoprotein | E. coli | 43 | glycosylated | Hershewe et al.135 |
| 2023 | PD glycosylated with ETEC O78 glycan | Glycoprotein | E. coli | ∼150 | glycosylated | Warfel et al.45 |
Abbreviations: scFv: antibody single-chain variable fragment, vtPA: variant of human tissue-type plasminogen activator, GM-CSF: granulocyte macrophage colony stimulating factor, IGF-I: insulin-like growth factor I, cIFN-α: consensus human interferon-alpha, rhGM-CSF: human granulocyte macrophage colony-stimulating factor, PD: Haemophilus influenzae protein D, IgG: Immunoglobulin G, β2AR: β2 adrenergic receptor, GPCR: G protein coupled receptor, HBc: Hepatitis B core antigen, HuNoV VP1-GII.3: Human norovirus genotype GII.3 VP1 capsid gene. *Indicates non-batch CFE reaction yield.
3.1.1. Proteins Requiring Folding Chaperones
Many proteins, particularly large, multi-domain proteins, require the assistance of chaperones372,373 and other regulatory mechanisms (e.g., modulated translation rates374) to properly fold. Interestingly, E. coli CFE systems typically have a translation rate more than an order of magnitude lower than living E. coli(179) (∼1 amino acid/second), which is closer to that of eukaryotic translation375 and may be beneficial for proteins that require more time for cotranslational folding.374 However, this alone may not be sufficient to support proper folding of every complex protein in CFE systems. Furthermore, purified CFE systems by design contain no chaperones and extract-based systems may or may not have sufficient levels of endogenous chaperones to adequately support proper folding, and thus CFE systems are frequently supplemented with additional chaperones (e.g., DnaK/DnaJ/GrpE, GroEL/GroES).357,368 In one case, a ribosome targeted fusion of the HSP 70 chaperone (BiP), which has a function analogous to that of DnaK, resulted in improved soluble yields of eukaryotic proteins expressed in an E. coli CFE system.376 In another example, supplementing FkpA and SlyD prolyl isomerases as well as the SkpA deaggregase improved the yields of soluble antibody.368 Artificial chaperone systems consisting of polysaccharide nanogels have also been shown to improve the soluble expression of proteins prone to aggregation.377
In addition to traditional chaperones, fusion proteins (e.g., maltose binding protein (MBP), thioredoxin (TRX), small ubiquitin-related modifier (SUMO), glutathione-S-transferase (GST)), which are thought to improve the folding and solubility of proteins, have also been utilized in CFE.95 Interestingly, proteases like the tobacco etch virus protease are active in some CFE systems, enabling production of protein fusions and subsequent cleavage of the fusion partner in a one-pot format.95 CFE has also been used to both express chaperones and subsequently assess their impact on folding of a target protein, a strategy that could accelerate researchers’ ability to identify the chaperones required for a given protein.378 Chaperone function and specificity can also be studied using CFE systems as has been done for the DnaK/DnaJ/GrpE chaperone system.379
3.1.2. Proteins Containing Disulfide Bonds
Disulfide bonds are critical elements of many proteins that serve to stabilize their structure.380 Disulfide bonds are formed in specialized compartments or organelles—the endoplasmic reticulum (ER) in eukaryotes and the periplasm in prokaryotes—which maintain an oxidizing environment to enable spontaneous disulfide bond formation.381 This oxidizing environment is enabled by systems of enzymes that maintain the redox potential and assist with the formation of the correct disulfide linkages.381 Both features of these compartments can be recapitulated in a CFE systems. An oxidizing environment can be maintained through the addition of small molecules (e.g., oxidized (GSSG) and reduced (GSH) glutathione).382 Extract-based CFE systems are also frequently treated with iodoacetamide (IAM), an alkylating reagent that inactivates the endogenous redox enzymes of the extract, to stabilize the redox environment.359 Furthermore, the enzymes responsible for disulfide bond isomerization, PDI in eukaryotes381 and DsbC in prokaryotes,381 can be supplemented to enable the proper formation of disulfide bonds.368
Many different proteins containing disulfide bonds have been successfully manufactured using CFE, including human hormones,363 cytokines,37,365 enzymes,359,360 and antibodies107,108,357,358,368,371,382 (Table 6). To quantify the progress of CFE systems capable of manufacturing disulfide bonded proteins, it is instructive to track the manufacture of antibodies.382 The first efforts to manufacture antibodies started with simpler single chain antibody variable fragments (scFvs) and yielded 8.3 μg/mL of protein.357 Modern CFE systems enable the manufacture of full-length heterotetrametric IgG antibodies with yields from 114 μg/mL (Chinese hamster ovary cell (CHO) CFE, glycosylated)107 to 150 μg/mL (N. tabacum BY-2 CFE extracts)99 to approximately 1,500 μg/mL (E. coli CFE, aglycosylated).368 These advances enable CFE not only to be a research tool for studying and prototyping disulfide bonded proteins, but also to be a viable manufacturing platform.37,43
3.1.3. Membrane Proteins
Despite the absence of an intact cellular membrane, CFE platforms offer ways to enable functional membrane-bound protein expression (Figure 7).383,384 During extract preparation, native cellular membranes are disrupted at the lysis step and self-assemble to form vesicles385 (e.g., cytoplasmic/periplasmic membrane vesicles in bacteria;135,386 ER/golgi microsomes in eukaryotic extracts387). In E. coli CFE, functional membrane-bound proteins (e.g., oligosaccharlytransferases) have been enriched in E. coli extracts prior to lysis,135,269 but synthesis of functional membrane proteins has frequently required supplementation of purified native membrane vesicles.385,386,388 Similarly, in eukaryotic CFE platforms, cell-free membrane protein expression has been demonstrated using both supplemented (i.e., exogenous microsomes in rabbit reticulocyte extracts389) and enriched (i.e., endogenous microsomes in HeLa390 and insect387,391,392 extracts) microsomes.383 While native vesicles preserve the membrane composition of the extract source strain they offer little control over the membranes in the final extract, and vesicle supplementation requires lengthy purifications.386
Figure 7.

Components for expressing soluble membrane proteins using CFE. Membrane mimics in the form of native vesicles, synthetic vesicles, nanodiscs, detergents, and oil droplets can be supplemented to or enriched in cell-free systems to enable the soluble expression of membrane proteins with various numbers and lengths of transmembrane domains (TMDs). Alternatively, cell-free systems have been used to express insoluble membrane proteins for subsequent refolding.
Synthetic membrane vesicles or liposomes can be supplemented to CFE reactions as alternatives to native membranes and enable precise control over membrane properties. Examples include aquaporin Z393 and ATP synthase in 1,2-dioleoyl-sn-glycero-3-phosphocholine (DOPC) liposomes.366 By changing properties of the lipids such as saturation, acyl chain length, charge, and curvature, or incorporating additives (e.g., polymers, cholesterol), membrane properties (e.g., elasticity) and membrane protein expression efficiency can be modulated.384 This was shown to improve folding of a mechanosensitive channel protein (MscL) in PURE.394 The hydrophobic matching of the protein and the lipid membrane thickness is also an important factor for expression and insertion into the membrane.395,396 Synthetic vesicles in PURE have also been used to systematically stop and start translation to study membrane protein folding and the effects of different lipid compositions.397,398 Due to the control over vesicle formation, membrane vesicles can also be used to encapsulate CFE reactions to make artificial cells (Section 3.9.3). Biomolecular condensates provide an alternative platform for synthetic cell formation, linking the chemical and cellular stages of living systems.399
Nanodiscs—noncovalent nanoscale assemblies composed of a lipid bilayer surrounded by amphiphilic scaffold proteins400—have also been used to synthesize diverse membrane proteins including oligosaccharyltransferases (OSTs),270,401 the methyltransferase Opi3,402 the antigenic vaccine carrier pore protein PorA,141 and the chemokine GPCR CCR5.384,403 In nanodiscs, both sides of the lipid membrane are exposed to the CFE reaction, which enables the study of both extracellular ligand binding and cytosolic activity of proteins without concern for orientation in the membrane.383 However, the design of nanodiscs (e.g., size) is more constrained than synthetic vesicles due to the need for membrane scaffold proteins. Nanodiscs remain an attractive membrane mimic as they are commercially available and can be affinity tagged for purification of an associated membrane protein.383,404
Other synthetic components have been used to express functional membrane proteins in CFE. For example, detergents—typically used to resolubilize protein aggregates by forming micelles with hydrophobic centers—have been supplemented into CFE for soluble membrane protein expression.270,362,405,406 Alternatively, precipitate-forming CFE takes advantage of the rapid and high levels of insoluble membrane protein expression followed by solubilization with detergents or reconstituted in membranes.407,408 A major disadvantage of detergents is their disruption of membranes and incompatibility with some downstream applications.383,409 Oil droplets have also been added to purified CFE to enable single pass transmembrane protein expression and assessment of surface receptor activity.410 Work enabling in situ lipid synthesis could enable production of lipids and membrane mimics in CFE itself toward more autonomous expression systems.411,412
3.1.4. Glycosylated Proteins
Protein glycosylation—the enzymatic process that attaches oligosaccharides to amino acid side chains in proteins—is found in all domains of life413 and is of interest for a variety of therapeutic applications.414,277 For example, N-linked protein glycosylation is present on many therapeutically relevant proteins, such as antibodies.415,416 Glycosylation occurs in the endoplasmic reticulum (ER) of eukaryotes where a multisubunit enzyme called an OST is responsible for the transfer of a lipid-linked oligosaccharide (LLO) to a protein at a specific amino acid residue recognized by the enzyme.415 Studying and manufacturing glycoproteins in cells, especially with homogeneous glycans, remains difficult, and CFE offers a new strategy to make such products.
Eukaryotic organisms possess post-translational glycosylation systems that have allowed eukaryotic extracts to be used to study glycosylation mechanisms417,418 and produce glycoproteins (e.g., antibodies,108 erythropoietin (EPO)419).28,420 Specifically, microsomes of ER and Golgi components can successfully glycosylate proteins containing signal sequences for translocation.177N-linked glycoproteins have been synthesized in multiple eukaryotic CFE systems including CHO, insect,391 human,68 hybridoma,109 and tobacco.99 Streamlined microsome enrichment protocols391 and semicontinuous reaction formats108,421 have increased glycoprotein yields.
While many prokaryotic organisms do not have endogenous glycosylation pathways,422 glycoengineering strategies have emerged to build glycoproteins in bacterial CFE systems. For example, E. coli CFE systems, which do not have the capability to glycosylate proteins natively,177 have been engineered to glycosylate proteins via two pathways: (i) OST-dependent - which uses an en bloc transfer mechanism and (ii) OST-independent–which harnesses individual glycosyltransferases to transfer each sugar and to build glycan structures from the bottom up. The first instance of cell-free glycoprotein synthesis was achieved with the supplementation of the well-characterized OST from Campylobacter jejuni (CjOST)423 and LLO424 from C. jejuni (Cj) to a CFE reaction to glycosylate a cell-free synthesized protein containing a glycosylation tag.425 A one-pot cell-free glycoprotein synthesis platform was then developed that expresses the membrane-bound glycosylation machinery in the extract source strain prior to extract preparation.269 By enriching exogenous glycosylation systems (e.g., CjOST and LLOs) in E. coli extracts or expressing them with CFE,270,401 glycoproteins can be produced.135,269,277,426 OST immobilization can also enable continuous glycoprotein synthesis in microfluidic devices.427 CjOST and homologues have been used to transfer the human N-linked trimannose core, C. jejuni and C. lari glycans, as well as various O-antigens from pathogenic bacteria.141,269 O-linked glycosylation has been shown for the C. jejuni glycan as well as human cancer-related mucin-type T and Tn antigens.135,428 CFE has also been used to rapidly synthesize low-cost and thermostable45,135 protective antibacterial conjugate vaccines against Francisella tularensis from lyophilized reactions.141
OST-independent glycosylation in E. coli CFE can enable high-throughput screening and building of glycan biosynthesis pathways.277,429 These methods have also been used for high-throughput screening of glycosylation tag sequence specificity429 and site preferences of N- and O- linked glycosyltransferases from bacteria and humans.430 This method also enabled the screening and identification of orthogonal enzymes capable of multisite glycosylation of target proteins.431 Another study on the modular expression and mixing of different glycosyltransferases and activated sugar donors in the CFE resulted in 20+ new glycan structures with possible applications for therapeutics, biomaterials, and antitoxins.277
3.1.5. Proteins With Other Post-Translational Modifications
Additional post-translational modifications (PTMs) are enabled in extracts by native biological machinery present in eukaryotic extract source strains. Phosphorylation,432,433 lipidation,434,435N-acetylation,436,437 ubiquitination,438 as well as signal peptide processing,391 have all been observed in cell-free systems derived from insect or rabbit reticulocyte cells. ER microsomes are required for many of these modifications.28,177 While E. coli extracts are not typically used for making PTMs, a prenyltransferase-enriched E. coli extract enabled higher yields of prenylated protein than reported in insect cell-free systems.439 Alternatively, ncAAs can be leveraged to add PTMs to proteins, such as the incorporation of L-phosphoserine into proteins cotranslationally in E. coli cell-free systems440 (Section 3.2).
3.1.6. Toxic Proteins and Peptides
The lack of a cell viability constraint makes CFE a promising alternative to manufacture proteins that are toxic to cells when expressed.441 As a result, CFE has been used for decades to study the activities of peptide and protein based toxins, and with advancements in reaction productivities, it is also now being considered as a viable manufacturing route for potential therapeutic molecules that are not possible to be produced in traditional expression systems.140 To date, many functionally active proteins with diverse toxic effects from organisms ranging from bacteriophages to eukaryotes have been expressed using both eukaryotic and prokaryotic CFE systems (Table 7). Even for toxic proteins with activities that impact transcription or translation, it may be possible to express the protein in CFE. For example, onconase—a nuclease that degrades tRNAs and has potential as a cancer therapeutic—was successfully manufactured in E. coli based CFE by supplementing additional tRNAs over the course of the reaction.140 It is important to note that not all CFE systems are suitable for a given target protein, particularly for small peptides,442 and thus it may be important to evaluate multiple different CFE systems for the production of the desired molecule.
Table 7. Examples of Functionally Active, Toxic Proteins and Peptides Manufactured Using Various CFE Systems.
| Year | Toxic Protein(s) | Toxic Protein Origin | CFPS System | Reference |
|---|---|---|---|---|
| 2001 | Cecropin P1 | Ascaris suum | E. coli | Martemyanov et al.264 |
| 2021 | Hemolysin BL | Bacillus cereus | Chinese hamster ovary (CHO) | Ramm et al.443 |
| 1974 | Colicin E1 | E. coli | E. coli | Eichenlaub444 |
| 2010 | α-Hemolysin | E. coli | E. coli | Chalmeau et al.445 |
| 2018 | Colicins Ia, E1, and E2 | E. coli | E. coli | Jin et al.446 |
| 2022 | Heat-labile enterotoxin (LT) | E. coli | Chinese hamster ovary (CHO) and Spodoptera frugiperda (Sf21) | Ramm et al.447 |
| 1998 | Enterocins L50A and L50B | Enterococcus faecium | E. coli | Cintas et al.448 |
| 2021 | U2-sicaritoxin-Sdo1a (venom component) | Hexophtalma dolichocephala | PURE, E. coli, and Spodoptera frugiperda (Sf21) | Lüddecke et al.442 |
| 2011 | Pierisin-1b | Pieris rapae | Spodoptera frugiperda | Orth et al.449 |
| 2016 | Exotoxin A | Pseudomonas aeruginosa | E. coli | Krinsky et al.450 |
| 2011 | A2 | Qubevirus durum | E. coli | Smith et al.451 |
| 2016 | Onconase | Rana pipiens | E. coli | Salehi et al.140 |
| 2022 | Cholera toxin (Ctx) | Vibrio cholerae | Chinese hamster ovary (CHO) and Spodoptera frugiperda (Sf21) | Ramm et al.447 |
| 2013 | Thermostable direct hemolysins | Vibrio parahemolyticus | E. coli | Bechlars et al.452 |
| 2017 | Thermostable direct hemolysins | Vibrio parahemolyticus | E. coli | Dondapati et al.453 |
3.2. Incorporation of Noncanonical Amino Acids into Proteins
Noncanonical amino acid (ncAA) incorporation provides a powerful opportunity to decorate proteins with user-defined chemistries. These chemistries are diverse,454,455 ranging from modifications of the side chain to changes that can alter the underlying polypeptide backbone, and they provide unique opportunities to study protein structure and expand protein function.456−458 The interface between ncAA incorporation and CFE provides several advantages. For instance, users can suppress codons in ways that would be technically difficult to accomplish in cells or use ncAAs that would interfere with cell viability. In addition, the open reaction environment allows for both the addition and removal of components that facilitate efficient ncAA incorporation, such as the inclusion of engineered aaRSs and tRNA pairs and the usage of ncAAs that do not easily cross the cell membrane. Thus, CFE provides an efficient reaction format for ncAA incorporation into peptides and proteins (Figure 8).
Figure 8.
Incorporation of ncAAs into proteins using CFE. (A) Process for ncAA incorporation using CFE. (B) Important components involved in ncAA incorporation in CFE.
There are several ways to incorporate ncAAs in vitro, including global suppression and site-specific incorporation. Global suppression, or residue-specific incorporation, takes advantage of natural aaRS promiscuity to incorporate ncAAs in place of canonical amino acids.459 By replacing a canonical amino acid with a close structural analog within a CFE reaction, an aaRS can acylate its cognate tRNAs with the structural analog to enable global suppression. This methodology efficiently recodes multiple sense codons for a ncAA but is limited by aaRS substrate specificity and requires residue-wide substitution of the ncAA. In site-specific incorporation, a codon is reprogrammed to encode a ncAA through a variety of techniques such as chemical modification of the acylated amino acid on the tRNA,460−462 chemoenzymatic ligation of an acylated adenosine onto a 3′ truncated tRNA,463−465 enzymatic acylation using aaRSs,466 and flexizyme-mediated methods.467 Of these methods, enzymatic acylation using aaRSs and flexizyme-mediated methods have become the most widespread. aaRSs are much better catalysts, enabling higher yields of acylated tRNA, but have limited substrate range.466 On the other hand, because flexizymes recognize the leaving group of an activated ncAA and the near-universally conserved 3′-CCA of a tRNA, they enable nearly limitless customization of the genetic code at the cost of the catalytic performance.467 Both options, however, enable more specific engineering of the genetic code compared to global suppression systems.
3.2.1. Incorporation of Noncanonical Amino Acids Using PURE
ncAA incorporation within the PURE system benefits from the flexibility in adding, removing, and titrating purified translation components. Many examples have used the Flexible in vitro Translation system (FIT), which couples flexizyme-based tRNA aminoacylation with PURE.468 In FIT, tRNAs charged with flexizymes are added into custom PURE reactions where competing translation components, such as amino acids or aaRSs, are omitted. This enables radical reprogramming of the genetic code. Additionally, FIT can be supplemented with EF-Tu and EF-P, in conjunction with tRNAs engineered to recruit those translation factors, to enable the ribosomal incorporation of backbone-modified ncAAs, such as D-, β-, cyclic β, cyclic γ, aminobenzoic acid, α-aminoxy, α-hydrazino, and thioamide-containing ncAAs.469−475 aaRSs can also be used in PURE to enable genetic code expansion. Orthogonal, amber-suppressing pyrrolysyl-tRNA synthetases and tRNA pairs like those from Methanosarcina mazei and Candidatus Methanomethylophilus alvus are widely used.476,477 Genetic code reprogramming by substituting endogenous aaRSs for more promiscuous mutants, such as PheRS A294G, LeuRS D345A, and editing deficient ValRS T222P, has also been reported.478,479
One application for the diversity of ncAAs compatible with PURE-based systems is peptide drug discovery. A method called Random nonstandard Peptide Integrated Discovery (RaPID) couples FIT with mRNA display to identify ncAA-containing peptides that have potent binding toward a protein of interest. ncAAs in RaPID can serve many purposes, such as a cyclization handle to form peptide macrocycles,511 as chemical warheads or drug delivery vehicles,512,513 or merely as a source of additional diversity in peptide libraries. Often, the ncAAs are indispensable for the function of interest. Several excellent reviews provide in-depth discussions about peptides discovered in RaPID.230,514,515 In addition, ncAA incorporation can also facilitate the discovery of peptide drugs that mimic ribosomally synthesized and post-translationally modified peptides (RiPPs).516 One benefit is the ability to enable chemistries that mimic those typically installed post-translationally by enzymes. For example, azoles and dehydroalanines can be installed into peptides by chemical modification of ncAAs, which expands the chemical diversity of peptides that can be ribosomally synthesized.517,518 These unnatural peptides can be modified with RiPP biosynthetic enzymes to further diversify peptide chemistries. The ncAAs incorporated through these methods thus may enable the discovery of novel peptide drugs.
PURE uniquely enables the study of genetic code architecture, as in vivo changes to the genetic code are largely lethal. For example, reconstitution of PURE with a minimized set of tRNAs, aaRSs, and amino acids, can vacate codon boxes to create a genetic code with only NNS (S = G or C) codons.519 PURE can be further simplified by using only 21 tRNAs, thus creating a minimal 20 codon genetic code.520 This should enable the newly vacated codons to be reassigned for ncAAs. In another example, the codons for serine, leucine, and alanine were swapped to create genetic codes that could only correctly decode functional proteins using a set of mis-acylated tRNAs.521 PURE might also enable studies on nonspecific decoding of codons through wobble-base pairing, which might highlight how new genetic codes might be used for more accurate ncAA incorporation.522
PURE provides an environment within which translation factors can be studied and engineered. PURE and flexizyme-charged tRNAs were used to elucidate identity elements for EF-P binding in Pro-tRNAPro.481,487 This study systematically dissected how the amino acid and tRNA sequence affect EF-P recruitment to alleviate ribosome stalling at proline codons. Similarly, studies have identified tRNA identity elements that serve to recruit EF-Tu, which helps to transport ncAA-charged tRNAs to the ribosome.485,488 Other engineered translation components in addition to tRNAs can be used within PURE. Ribosomes that recognize tRNAs containing non −CCA 3′ ends have been engineered to enable two orthogonal genetic codes to function in parallel, allowing for two different peptides to be synthesized from the same DNA template.523 These technologies might allow for the in vitro engineering of ribosomes within PURE to polymerize nonpeptidyl backbones.
PURE has proven itself as a uniquely powerful technology that enables research into areas that are inaccessible by lysate-based CFE or cell-based methods. However, it may not be appropriate for all applications in this space. For example, average protein yields from PURE are typically 5–10x less than crude-extract based systems.21,23 This may limit the ability to synthesize proteins at the scales required for characterization studies. Additionally, because each translation component in PURE must be purified, PURE is significantly more expensive and labor intensive than extract-based approaches ($1,000 per mL reaction compared to $4.67 per mL in batch mode, respectively).1 Despite these disadvantages, PURE is positioned to enhance our understanding of how translation can be engineered for ncAA incorporation.
3.2.2. Incorporation of Noncanonical Amino Acids Using Extract-Based CFE
Cell extracts provide unique features for ncAA incorporation into proteins. They are higher yielding and simpler to prepare than PURE, but the presence of cytoplasmic components (e.g., proteases; nucleases; endogenous aaRSs, tRNAs, and amino acids) can be detrimental for ncAA incorporation into proteins (Table 8). Recent progress has shown that cells can be engineered to produce extracts that are more effective for ncAA incorporation into proteins. For example, to combat premature protein truncation by Release Factor 1 (RF1), which decreases amber codon suppression efficiency,497,524 extracts can be prepared from cells either containing a thermosensitive RF1524 or lacking RF1 completely.525,526 Two E. coli derivatives where RF1 is removed from the genome, the C.321ΔA and B.95ΔA strains, can be used for extract preparation. Further genome engineering of C.321ΔA to remove negative effectors of CFE (e.g., proteases, nucleases) and to introduce a T7 RNA Polymerase yielded the 759.T7 strain, which can incorporate two p-acetylphenylalanine (pAcF) residues into a single sfGFP at yields >2 g/L and 40 instances of pAcF into an elastin-like polypeptide in a batch reaction.23,117 Extracts made from B.95ΔA are also highly productive, showing >3 g/L into 1TAG-sfGFP in a continuous-exchange format. These engineered extracts are thus highly useful tools for ncAA incorporation in CFE.527
Table 8. In Vitro Translation Reactions Can Be Customized by the Addition and Removal of Translation Components to Facilitate ncAA Incorporationa.
| Translation Component | PURE CFE | Exctract-Based CFE |
|---|---|---|
| tRNAs | + tRNAfMet478,480−482 | + tRNAPyl489,490 |
| + tRNAAsnE2,483,484 variants of tRNAAsnE2485 | + tRNAM.jannashii Tyr491 | |
| + tRNAGluE2486 | - RNase digestion of tRNAs492,493 | |
| + tRNAPro1E2487,488 | - Chromatographic depletion494,495 | |
| + tRNAPhe | -Inactivation using antisense oligonucleotides496 | |
| + tRNALeu | ||
| + tRNAPyl | ||
| aaRS | + PheRS A294G E. coli(478) | + M. jannaschii TyrRS variants117,497−500 |
| + LeuRS D345A E. coli(478) | + M. alvus PylRS and variants thereof501 | |
| + ValRS T222P E. coli(479) | + M. mazei PylRS and variants489,490,502 | |
| + PylRS M. mazei(476,477) | ||
| + PylRS M. alvus(476,477) | ||
| - Endogenous aaRSs through exclusion in PURE dependent on choice of codon and amino acid.468 | ||
| Elongation Factors | + EF-Tu and variants503−505 | + EF-Tu and variants440,506 |
| + EF-P470,488 | ||
| Release Factors | - RF-126 | - Inhibition of RF-1 |
| - Degradation of RF-1507 | ||
| Ribosome | + 040329508 | - Depletion of ribosomes through ultracentrifugation509 |
| + in vitro constructed ribosomes support ncAA incorporation510 |
Supplemented components are indicated with a (+) and withheld or removed components are indicated with a (−).
CFE reaction formulations themselves can also be customized for efficient ncAA incorporation. First, translation components such as the aaRS and tRNA can be coexpressed in CFE along with the protein of interest.491,502,528 This strategy diverts resources away from synthesizing the protein of interest but can be useful for applications such as aaRS or tRNA engineering. Second, heterologous translation components can be enriched in cell extracts by overexpression in cells prior to lysis.489 This is particularly useful with components that are difficult to purify, such as the commonly used M. mazei and M. barkeri pylRSs. Finally, endogenous translation components can be removed from cell extracts. For example, endogenous E. coli tRNAs can be inactivated or depleted using chromatography,494,495 enzymatic digestions,492,493 and antisense oligonucleotides,496 and endogenous amino acids can be dialyzed away to enable the use of global incorporation strategies,529,530 notably for the incorporation of isotopically labeled amino acids as spectroscopic probes.46,531 These examples, as well as others (Table 8), show strategies in CFE to improve, simplify, or expand the range of ncAA incorporation.
ncAA-containing proteins made using CFE have many applications. For example, CFE has been used to commercially synthesize antibody-drug conjugates that incorporate ncAAs such as p-azidophenylalanine and p-azidomethylphenylalanine532 that serve as click chemistry handles. ncAAs may also be incorporated into enzymes, enabling the studies on how novel chemistries could improve enzyme catalysis and introduce enzyme regulation strategies.533,534 Conversely, the incorporation of ncAA mimics of post-translational modifications (PTMs) can inform the function of PTMs.535 A CFE system engineered for efficient phosphoserine incorporation, for example, was used to study the role of phosphorylation patterns on human MEK1 kinase activity.440 In addition, ncAAs can serve as biophysical probes for spectroscopic studies, including fluorinated ncAAs for 19F NMR536 and cobalt-binding ncAAs as paramagnetic labels,537 and as fluorescent probes to detect changes in local environment538,539 and to detect long-range conformation changes using FRET.540−542 Thus, CFE has enabled the incorporation of many ncAAs that serve numerous applications.
Although both PURE and extract-based CFE systems have their merits for ncAA incorporation into proteins, the trade-off between yield, cost, genetic code reprogramming capability, and efficiency must be carefully considered.
3.3. Scale-Up for Decentralized Protein Manufacturing
CFE systems are poised to enable a new approach to decentralized manufacturing.141,543,544 To enable this vision, cell-free systems are scaled beyond their typical μL laboratory-scale volumes. Over the past 15 years, technological breakthroughs have shown that CFE reactions are linearly scalable between the μL and 100 L scales, enabling high yields of proteins.37,545 For example, in some initial studies, reaction volumes of 2 mL, 50 mL, and 1 L all produced human insulin-like growth factor-1 (IGF-1) at the same rate in stirred tank reactors.546 Beyond protein synthesis, cell-free reactions that require in vitro enzymatic activity but not protein synthesis547 have been scaled to 20,000-L reactors to produce myo-inositol with heat-purified enzymes.548
A few key works highlight important considerations for scalable, cell-free protein production.37,71 First, cost remains a major consideration for large scale reactions. The primary cost determinants are lysate preparation, carbon and energy sources in the reaction, DNA, and additives such as CoA, NAD, and NTPs.1,45,157 As discussed in Section 2.3.1, high-cost phosphorylated energy substrates such as PEP can be replaced by more economical substrates such as glutamate,157 which harness oxidative phosphorylation and native metabolism of the cell extract. Second, extract processing must also be optimized for efficient large-scale production. Source strain modifications can improve yields by stabilization of extract components such as amino acids.549 To increase process throughput, disk stack centrifugation has been used to process hundreds of liters of cells for extract production and can be combined with increased dilution of the cell suspension before lysis to improve extract clarification.37 Extract processing can also be streamlined by removal of the dialysis step,128,130 however it should be noted that dialysis has been identified as important for applications involving transcription from native promoters using RNA polymerases present in the extract.122 Third, transport limitations must be considered as oxygen is required for reaction formulations relying on extract metabolism for energy regeneration. One scale-up issue is that CFE reactions can foam when agitated in traditional bioreactors. The addition of antifoam agents has been optimized to reduce foaming without negative impacts on yields.546 While most optimization has focused on bacterial systems, recent work in a N. tabacum BY-2 CFE system highlights that scale up with eukaryotic systems is also feasible.71
Amenability to lyophilization550 and reaction modularity make CFE an ideal approach for decentralized biomanufacturing. For example, point-of-care diagnostics142,143,250,551 and therapeutics139−141 can readily be distributed when lyophilized without cold-chain storage. Further, the Bio-MOD (biologically derived medicines on demand) platform—an integrated suitcase-sized device for small-scale GMP production and purification of therapeutic proteins such as GCSF and erythropoietin—was developed using CFE from lyophilized CHO cell extracts.543 Thermostability of CFE reactions can be improved through the addition of lyoprotectant additives or strategic separated storage of reaction components;45,138,147,148,262,552 these strategies have resulted in the cell-free production of pyocin,138 nanobodies,553 a protein-based vaccine (DT),553 conjugate vaccines,45,554 and a cancer therapeutic (ErA) following storage above ambient conditions.552
Numerous works have sought to decrease cost of CFE reaction components while maintaining high protein yields.41,43,45,155,156 However, reported cost analyses of CFE reactions at the laboratory scale often do not account for costs associated with bulk reagent purchase, labor, or capital equipment.45 Looking forward, we anticipate that continued investment in biomanufacturing leveraging CFE systems will enable additional large scale economic analyses555 and the more accurate prediction of cell-free protein production costs at a variety of scales. Beyond cost, we anticipate that innovations in purification and downstream quality control processes from lyophilized cell-free expression reactions will enable scaled batch production and distribution of protein products at the point-of-need.
3.4. CFE for Measuring and Engineering Macromolecular Interactions
CFE systems have been used to study and engineer macromolecular interactions involving proteins, nucleic acids, lipids, and small molecules. The open nature and high-throughput potential of CFE systems makes them ideal for probing and engineering these interactions. When working with individual proteins, the features of CFE systems allow for a nearly order of magnitude improvement in speed and throughput when compared to equivalent cell-based processes that are bottlenecked by cloning, transformation, and cell culture.116,282,556 When working with libraries of proteins, CFE systems offer a variety of ways to couple genotype to phenotype and can achieve library sizes several orders of magnitude greater than their cell-based counterparts.339,340 Here we summarize the work done to measure and engineer molecular interactions, with a focus on protein–protein interactions (PPIs).
3.4.1. High-Throughput Screening of Macromolecular Interactions
Several standard methodologies have been used to quantify PPIs in the complex CFE environment without purification in microplates (Figure 9). Protein complementation assays,557 where reporter proteins are fragmented and genetically fused to the proteins of interest, have been used with CFE to probe natural PPIs,558,559 identify inhibitors of PPIs,560,561 and prototype PPI-based logic gates.263 Fluorescence correlation spectroscopy, which uses confocal microscopy to observe fluorescently labeled molecules that transit through an observation volume to measure the oligomeric state of the protein(s) (Figure 9),70,562 can evaluate the interactions and oligomeric state of the expressed proteins or complexes directly in a CFE reaction.70,556,562−565 Enzyme-linked immunosorbent assays (ELISAs),566 a gold standard method in PPI screening relying on immobilization of antigens on a surface and detection with an enzyme- or fluorophore-linked antibody, have been utilized to screen antibodies produced in CFE.188,567,568 The amplified luminescent proximity homogeneous linked immunosorbent assays (AlphaLISA),569 an in-solution ELISA-like assay,569 tolerates several crude CFE systems186,556,570 and is well suited for high-throughput screening because it does not require wash steps. AlphaLISA has been utilized to measure the interaction specificities between several natural proteins,556,563−565 profile mouse genes for autoantigenicity,570 as well as evaluate antibodies116 and computationally designed binding proteins.189 An in vitro two-hybrid (IVT2H) assay, where PPIs are quantified by transcriptional activation of a reporter gene, has also been developed for CFE.571 Collectively, these measurement techniques enable researchers to probe hundreds to thousands of PPIs using CFE in microplates.
Figure 9.
Methods for evaluating PPIs in CFE that do not require purification of the interaction partners.
CFE-based self-assembling microarray assays (Section 2.7) can be used to measure PPIs by exposing the microarray to a labeled target protein, where interactions are measured by the intensity of the label measured at each microarray spot. Microarrays have been used to map the interaction network of 841 interactions between 29 human replication proteins,295 profile antibody target antigens,572−574 investigate protein small molecule interactions, and more.575,576
Microfluidic systems are also commonly employed to measure PPIs. The protein interaction network generator (PING) is a microfluidic system that combines CFE with the mechanical trapping of molecular interactions (MITOMI) to evaluate the interactions of the expressed proteins.307 The system was used to evaluate 1,849 interactions between 43 Streptococcus pneumoniae proteins in quadruplicate, mapping a rich interaction network. PING also enables the evaluation of membrane proteins, with iterations of this technology enabling the simultaneous profiling of 2,100 membrane proteins to characterize host–pathogen interactions.577 A similar platform has also been used to study and engineer protein–DNA interactions.303,306 Droplet-based microfluidic systems have further been used to evolve peptide binders using the IVT2H assay.578
3.4.2. In Vitro Display Technologies
CFE has the unique benefit of enabling researchers to create a direct physical connection between a gene’s transcript and its protein product, generating a macromolecular complex linking genotype and phenotype. These approaches, typically referred to as in vitro display technologies, have several benefits for both the measuring and engineering the properties of the displayed proteins over cell-based counterparts.336 Library size in CFE systems is limited by number of ribosomes and DNA or RNA templates present as opposed to the transformation efficiency, enabling library sizes of up to 1014,339,340 appreciably larger than the typical maximum library size of 109 to 1010 in phage- or cell-based display systems.338,579 Library assembly and diversification steps can also be performed without transformation and cell growth, leading to faster selection cycles. Furthermore, the nonliving nature of CFE reactions enables methods for creating genotype to phenotype linkages that would not be possible in living systems. Taken together, these features have led to the development of several different in vitro display technologies.
Ribosome display maintains the genotype to phenotype linkage by stalling the ribosome on its transcript without releasing the polypeptide chain from its acceptor tRNA337,580−582 creating a ternary complex that can be used to select for or measure different properties of the displayed protein.337 The ribosome is stalled on the transcript through the omission of a stop codon or a peptide-based stalling sequence337,580 (Figure 10). Ribosome display has been used to engineer to a variety of different binding proteins, including single chain variable fragments (scFvs),583,584 antigen binding fragments (Fabs),585 designed ankyrin repeat proteins (DARPINs),586 and single domain antibodies.587−590 Of note, one of the strongest affinity antibody–antigen interactions reported to date was engineered using ribosome display using extended off rate selections for affinity maturation.583,591,592 Ribosome display has also been used to select for improved protein stability using chemical denaturants345 and protease treatment.346 A limitation of ribosome display is that experiments must be carried out under conditions where the template mRNA is not subject to degradation and the ribosome ternary complex is stable. However, it should be noted that comparisons of display technologies commonly overlook reports that the ternary complex can be stable even under mild denaturing stresses including elevated temperatures.580,593
Figure 10.

Schemes for ribosome and mRNA display.
mRNA display establishes a genotype to phenotype linkage by covalently linking the nascent peptide chain to the mRNA using an mRNA-puromycin conjugate338,340,350,594,595 (Figure 10). Puromycin is an antibiotic that interferes with translation by binding to the ribosomal A site and accepting the nascent peptide chain.338 By conjugating puromycin to the mRNA transcript, a covalently linked peptide-mRNA fusion can be generated. mRNA display has been used widely for the selection of binding proteins from random linear peptides,596 antibody mimics,597 single domain antibodies,598 and antibody antigen binding fragments.599,600 Beyond simple binding selections, various versions of mRNA display have also been used for binding selections against whole cell targets,601 measurements of functional proteomics,341−343 measurements of binding kinetics,602 selections for enzyme activity,349,350,603 selections for stability using chemical denaturants,347,348 temperature,349 and proteases,604 as well as selections to probe the functional areas of random sequence space.605 mRNA display and derivatives thereof also feature prominently in the RAPID system230,514 (Section 3.2.1) where they have been used to select linear and cyclic peptides containing noncanonical amino acids with properties akin to natural products.230,514 Limitations of traditional mRNA display include the stability of mRNA molecule itself as well as the effort required to prepare mRNA-puromycin conjugates. Several modifications have been made to mRNA display methods in order to address these limitations, including the use of cDNA display,606,607 transcription-translation coupled with association of puromycin linker (TRAP) display,608 cDNA TRAP display,609 and click display.610 These methods enable displayed proteins coupled to cDNA instead of mRNA,606,607,609,610 rapid and simple puromycin conjugation procedures,606−610 and one pot transcription and translation reaction formats.608−610
Beyond ribosome and mRNA display, many other CFE-based in vitro display techniques have been devised. CIS display relies on the ability of the RepA protein to bind the DNA template from which it was expressed, which has been used to select protein binders.611 Similarly, covalent antibody display (CAD) leverages the P2A enzyme to covalently link the expressed protein to its DNA template, enabling selection of scFvs.612 The STABLE method utilizes emulsions and streptavidin fused polypeptides to link the biotinylated template DNA to the expressed protein.613 Similarly, covalent DNA display614 and SNAP display615,616 use compartmentalization in emulsions and a fusion of the displayed protein to an enzyme that covalently reacts with a suicide inhibitor conjugated to the DNA template. Several different microbead display systems have been developed,617−619 all of which utilize in vitro compartmentalization in emulsions to immobilize both the CFE-expressed protein and its coding DNA template to a microbead that can be used to perform selections. Liposome display couples genotype and phenotype via compartmentalization using liposomes to enable the engineering of membrane proteins, toxic pore-forming proteins, transporters, and receptors.620
CFE display technologies have also been combined with modified next-generation sequencing technologies for high-throughput experimentation. Single molecule interaction sequencing (SMIseq)344 coupled with ribosome display was used to profile the specificity of a library of 200 scFvs against 55 different human proteins constituting 11,000 possible interactions.344 Protein display on a massively parallel array (Prot-Map)319 uses ribosome display coupled with fluorescently labeled target molecules to quantitatively assay binding, which was used to probe the sequence specificity of the M2 anti-FLAG antibody to thousands of possible antigens as well as to study the fitness landscape of more than 100,000 variants of the SNAP-tag self-labeling enzyme.319 In a similar approach dubbed deep screening, ribosome display and a sequencing flow cell were leveraged to quantitatively measure antibody binding affinity for millions of variants simultaneously to generate high quality data for training machine learning models.320 These techniques highlight the utility of CFE-based display technologies for directed evolution and performing high-throughput functional measurements.
3.5. CFE-Based Biosensors
Biosensors are analytical or diagnostic tools that leverage biological sensing elements (e.g., protein, antibody, nucleic acid, cell) to detect or quantify analytes as a visual or electrical signal output. Historically, whole-cell biosensors have been developed for metabolic pathway engineering, environmental detection,621−623 and biomedical analysis.621,624 However, the use of live cells for sensing provides a host of barriers for field-deployable implementation such as membrane permeability limitations, genetic instability, and biocontainment concerns.625−627 CFE-based biosensors allow for a full use of resources toward production of sensor or reporter machinery making them a compelling option over whole cell biosensors.1,262,626,628 Reaction conditions, like temperature629 and pH,630,631 can also be precisely controlled.
CFE biosensors can detect a range of analytes (e.g., small molecules, metals, nucleic acids, proteins) and have been recently reviewed in detail elsewhere.623 Cell-free biosensors can be lyophilized and rehydrated from both tube and paper-based formats142,249,262,632−634 with minimal loss in performance, allowing for long-term storage and application at the point-of-use. Notably, CFE biosensors hold promise to become low-cost, clinically relevant diagnostics as they can also remain active in complex sample matrices (e.g., serum,551,635 saliva,636,637 and urine638).
Quality biosensor design hinges on selecting a sensor and reporter system that suits the analyte and context it will be sensed in. Most CFE biosensors utilize allosteric transcription factors (aTFs) or aptamers/riboswitches to detect small molecules638 and ions143,639 (Figure 11). Nucleic acid detection is typically accomplished with CRISPR-Cas systems and/or toehold switches.250,640−642 CFE sensors can also be designed for the detection of missing reaction mixture components, such as amino acids.636,643−647 Recent work has further shown new detection strategies for antibodies648 and methyltransferases649 using nucleic acid probes.
Figure 11.

Mechanisms for common sensing elements of cell-free biosensors. (A) Paper and tube-based CFE reactions can be lyophilized for point-of-use application. (B) Common mechanisms of activator and repressor aTFs. Activators activate gene expression upon binding to the operator site and repressors repress gene expression. The aTF-operator binding mechanism is modulated by analyte binding, with analyte-binding either activating or inhibiting aTF-operator binding. (C) Example mechanism of fluoride riboswitch.143 The riboswitch folds cotranscriptionally into one of two mutually exclusive structures, conditional to fluoride binding. In the presence of fluoride, a pseudoknot enables expression of a reporter. In the absence of fluoride, a terminating hairpin represses expression. (D) Toehold switch mechanism. Toehold switches sequester the RBS and start codon RNA into a hairpin loop that is inaccessible to the ribosome. Upon viral trigger RNA binding to the upstream “toehold”, the hairpin loop opens, giving the ribosome access to initiate translation.
Biosensor readouts are typically either optical or electrochemical. Optical reporters have a light-based readout that can be fluorescent (e.g., GFP650−652), colorimetric (e.g., β-galactosidase (lacZ)249,653,654), or luminescent (e.g., luciferase637). Fluorescent and colorimetric readouts are well suited for point-of-use applications due to their simplicity and equipment-less readout. Electrochemical readouts, on the other hand, offer potential in their ability to interface with materials and equipment as portable diagnostics.655−659 For example, a cell-free biosensing reaction layered over a custom chip with microelectrodes provided a ligand-dependent voltage readout.656 Other works have shown that commercially sold personal glucose meters can be co-opted for CFE biosensors. To do this, researchers leveraged the electrons generated during the enzymatic oxidation of glucose as a readout for the detection of amino acids in extract-based CFE systems658 and viruses using purified protein-based systems or PURE.660 Below we detail the sensing elements used in CFE diagnostics.
3.5.1. Allosteric Transcription Factors
Allosteric transcription factors (aTFs) are proteins that bind an analyte (e.g., organic and inorganic small molecules, nucleic acids, proteins) and in turn regulate transcription of a gene downstream of an aTF-specific DNA sequence in the promoter region. aTF-based CFE biosensors have primarily been developed for the purpose of field-deployable small molecule detection, such as for environmental contaminants (e.g., pesticides142,661 and heavy metals262,630,639,662) and human health markers (e.g., drugs629,638 and metabolites638).
aTFs have been used with CFE systems to detect mercury,629,630,663−666 cyanuric acid,142 and other analytes.262,667,668 In these cases, previously identified and optimized aTFs have been tuned in CFE to achieve acceptable sensor performance.44 The aTF is typically supplemented to the sensing reaction in one of three ways: as a purified component;669 produced either through a separate CFE reaction that is then mixed with a fresh reaction containing reporter elements or concurrently with the reporter elements;122,639 or enriched in the extract during cell culture of the extract source strain,661 which can be mixed with nonenriched extract to achieve the desired sensing results. Further adaptation of an aTF system as a CFE biosensor often involves promoter optimization.142,661,666 For instance, promoters for the activator AtzR derived from cellular experiments had background expression in CFE in the absence of analyte and needed to be reoptimized in the cell-free environment.142 Other factors such as sensor/reporter concentration142,143,638,664,670 and reporter type143,665 serve as optimization handles for cell-free biosensors. The space of detectable molecules has been broadened beyond analytes that have known aTFs by enzymatically converting an analyte of interest into one with a defined aTF sensor, dubbed metabolic biosensing.638,661
3.5.2. Aptamers and Riboswitches
Aptamers and riboswitches are functional nucleic acids that can serve as biological sensing elements seen across all domains of life (Figure 11C).671−673 Specifically, aptamers are nucleic acids that are capable of binding to specific target molecules674 and riboswitches are RNAs that use an aptamer to bind to an analyte and regulate gene expression in response to that analyte.672 A number of riboswitches have been demonstrated in CFE systems and have been reviewed thoroughly elsewhere.675 Notable applications of CFE riboswitches have been toward artificial cell communication (e.g., theophylline676−678) and the detection of molecules relevant to human and environmental health (e.g., thrombin,679,680 fluoride,143,681 proteins,682 tetracycline,683 histamine,684 ciproflaxcin,684 and dopamine682).
Aptamers are appealing sensor elements because they can be easily synthesized via PCR, engineered in large libraries, and computationally designed.685,686 Most de novo developed aptamer and analyte binding pairs are engineered to target proteins and small molecules, although metal ions, cells, other nucleic acids, peptides, and carbohydrates have all been targeted as well.671 The main method for evolving de novo aptamers is using Systematic Evolution of Ligands by EXponential enrichment (SELEX), wherein a target molecule is incubated with a library of 1014 ∼ 1016 aptamer candidates.685 In conjunction with library screening, riboswitches can be designed in silico using the thermodynamic and kinetic principles that dictate nucleic acid folding.687 This was demonstrated for the design of protein-sensing riboswitches in CFE systems.682 Aptamer engineering methods specific to CFE-biosensors have been developed, such as using water-in-oil droplet sorting in microfluidic chips.684 In silico methods have also been used to design CFE-riboswitches for detecting proteins682 and human biomarkers.682
A salient demonstration of the potential for riboswitch sensors applied in CFE was the detection of fluoride using a riboswitch controlling expression of an enzymatic colorimetric reporter.143 This sensor system was lyophilized in microtubes, and reactions were able to sense fluoride at a limit of detection set by the EPA in the lab. To assess the robustness of this diagnostic in the field, diagnostics were deployed to rural Kenya and tested in the hands of nonexperts. The authors found that out of 57 tests, 89.5% of water samples were correctly classified under field conditions.688 This work was expanded upon by encapsulating the fluoride riboswitch in a lipid vesicle.681 The sensor’s access to analyte and protection from inhibitory enzymes were both controlled by altering the membrane permeability, expanding our repertoire of acceptable sample matrices, highlighting the potential for aptamers applied to cell-free systems.
3.5.3. Nucleic Acid Biosensors
Field-deployable nucleic acid biosensors are a critical need in modern medicine for the rapid and accessible detection of disease.689 CFE-based nucleic acid biosensors have enabled pathogen detection primarily through leveraging toehold switch sensing elements. RNA-based toehold switches detect specific pathogenic RNA-sequences through a regulating translation. In the absence of pathogen-specific trigger RNA, they sequester the RBS into a hairpin loop that blocks translation by the ribosome. Upon the sequence-specific binding of trigger RNA, the loop unfolds to allow translation and reporter signal.
Extract-based CFE biosensors have been developed to detect a number of global-health related viruses (e.g., SARS-CoV-2,257,637,652,659,690,691 respiratory syncytial virus,692 Norovirus,693 Ebola,249,257 Zika,250 and hepatitis694), bacterial markers (e.g., typhoid and paratyphoid,659 antibiotic resistance,659 gut microbiota695), and agriculturally relevant viruses (e.g., Potato Virus Y653 and cucumber mosaic virus696). These sensing reactions were all stably lyophilized and are estimated to cost $0.10-$2 per test.249,250,693,695
Toehold switches can be rapidly de novo designed to detect pathogen-specific trigger RNA sequences through well-defined and predictable Watson-Crick base pairing.697,698 Often, toehold switch sensor reactions are preceded by an isothermal amplification step on the sample,699−702 which enables the amplification of trigger RNA, the conversion of DNA to RNA triggers, or the addition of synthetic sequences to triggers. This has been frequently used to achieve improved detection limits in CFE diagnostics.250,659,690−693,695 CFE nucleic acid biosensor orthogonality can be further optimized by coupling a toehold switch with CRISPR Cas proteins, which was demonstrated for the discrimination between the African and American Zika strains with single-base resolution.250 To do this, the authors identified a naturally evolved single nucleotide polymorphism between the two strains that created a strain-specific PAM site for Cas9 cleavage in only the American ZIKV strain. In a pooled sample of both American and African ZIKV, the authors used isothermal amplification to append a synthetic trigger RNA sequence to the purified viral RNA upstream of the potential Cas9 cleavage site. In the American ZIKV strain, Cas9 cleavage produces a truncated, inactive trigger sequence. Therefore, the sensor only turns on for the African ZIKV strain, which produces full-length trigger RNA in the absence of Cas9 cleavage.
3.6. Genetic Parts and Circuit Prototyping
Characterizing the genetic parts (e.g., promoters, RBSs, terminators) necessary to predictably engineer biology can be laborious with engineering cycles spanning days to months depending on the host organism. Using CFE, large libraries of genetic parts can be quickly generated, assembled with a reporter gene into a transcriptional unit, and assayed at once by running CFE reactions in a variety of formats and volumes.333,703,704 Genetic parts can be assessed in a 96- or 384-microplate with a fluorescent protein reporter as a readout for relative activity.112,211,705−707 Higher throughput can be achieved by encapsulating CFE reactions in droplets,708 where gene variants can be coupled to fluorescent reporters so that fluorescence-activated droplet sorting (FADS) can be used for enrichment, accessing library sizes of >107 candidates per day.324,328,709,710 Multiplex characterization of 5′ regulatory sequences has also been demonstrated for pooled CFE reactions using RNA-seq as a readout.56
CFE has shown value for prototyping genetic parts from nonmodel organisms. Many nonmodel organisms are emerging as promising commercial chassis strains but are often challenging to engineer due to slow growth rates and limited tools for genetic manipulation. CFE systems utilizing extracts from nonmodel organisms (Section 2.1) sidestep these challenges and make it possible to prototype genetic parts in the context of their native host organism’s biological machinery. Using this strategy, genetic parts for Clostridium autoethanogenum, Bacillus megaterium, Streptomyces, Pseudomonas putida, and other organisms have been evaluated using their corresponding CFE system.51,54,56,58,91,112 Ensuring that prototyping results obtained in a CFE system still hold in the context of the living organism is important. To date, clear correlations have been observed between activity in CFE systems and in vivo.56,211,706,707 However, it should be noted that prototyping results can be influenced by the type of DNA template (Section 2.3.5)185,211 and the supplemented reaction components.711
Genetic circuits can be constructed from a combination of distinct genetic regulatory elements (e.g., DNA, RNA, proteins) that interact to carry out a given program or function (Figure 12). Although genetic circuits offer exciting possibilities for programming biological behavior,712 efforts to implement circuits in vivo often yield unpredictable results.261 CFE characterization could be beneficial, as circuit components and inputs can be easily tuned such as by adjusting the amount of DNA added to the open reaction environment. Basic circuit motifs have been implemented in CFE systems by combining regulatory genetic part elements together. AND-gates constructed from small transcriptional activator RNAs (STARs) and toehold switches713 have been demonstrated along with protein-based cascades, utilizing E. coli sigma factors and viral polymerases42,44,714 and RNA-based cascades.707,715 Through the combination of these elementary motifs, more complex logic can be achieved such as different types of feed forward loops,329,715−719 negative autoregulation720 and oscillators.721−723 Circuits implemented in CFE and in cells yield largely comparable results, suggesting that CFE can serve as a prototyping platform.707,713,718 However, like individual genetic parts, results can be convoluted by similar factors mentioned above as well as resource competition706 and batch to batch variability.724 Beyond prototyping circuits for cellular applications, CFE can also be used to study genetic circuits that would be challenging or impossible inside of a cell, an excellent example being the recent implementation of a biological circuit operating on a single DNA molecule in a fabricated biochip.311
Figure 12.
Components of genetic circuits in CFE.
3.7. Enzyme Screening
Enzymes are catalysts that carry out precise chemical transformations. As enzymes have become increasingly important as catalysts for chemical synthesis due to their speed, selectivity for substrates, compatibility with mild physiological conditions (aqueous, pH 7, 37 °C), ability to be assembled into enzymatic cascades, and evolvability, there is also an increasing need for technologies to support enzyme discovery, screening, and evolution.725,726
Large libraries of enzymes can be synthesized and screened using linear expression templates in CFE, bypassing traditional time and labor bottlenecks in DNA cloning. These enzymes are often produced in sufficient yield for multiple assays, which enables reactions to be scaled down to μL volumes and used in high-throughput methods (Section 2.7). The open reaction environment of cell-free systems enables enzymes to be tested without the barrier of the cell membrane and, depending on the application, without purification.727−729 CFE also enables the production of enzymes that may be otherwise toxic to life, due to the in vitro nature of the method (Section 3.1.6). Finally, enzymes can be tested within the context of active metabolism, which informs in vivo metabolic engineering applications, and provides access to native energy, cofactor, and substrate pools.730,731
A common method for studying and engineering enzymes in cell-free systems is to use well-by-well screens. In these methods, enzyme variants are encoded by DNA templates that can be generated by a variety of means, including from commercial gene synthesis, error-prone PCR, and saturation mutagenesis. Enzymes such as [FeFe] hydrogenases and aminoacyl-tRNA synthetases are amenable to CFE-based screening and engineering using these approaches.532,732 Droplet microfluidics have also been used to increase enzyme screening throughput (Section 2.7). For example, alkaline phosphatases were engineered to have >20-fold improved activity.318,733 Furthermore, microfluidic chips are capable of synthesizing, immobilizing, washing, and characterizing the kinetic parameters of >1,500 enzyme variants simultaneously.310 CFE technologies are poised to enable researchers to study and engineer enzymes at depths not previously possible.
Another method for screening and selecting enzymes with new properties using CFE is to compartmentalize reactions to link genotype and phenotype. In a proof-of-concept experiment, water-in-oil emulsions encapsulating E. coli S30 extracts that synthesized the methyltransferase from M. HaeIII could be selected from those synthesizing another protein by using resistance against restriction enzyme digestion.323 Water-in-oil emulsions are generally compatible with fluorescence-based droplet sorters or microfluidic droplet sorters, enabling large libraries to be screened rapidly.
In theory, any enzyme can be evolved using emulsions provided an appropriate substrate that does not diffuse through the emulsion and the desired chemical reaction can be observed for sorting or selection. A variety of enzymes, including β-galactosidase,734 penicillin G acylase,735 horseradish peroxidase,736 cellulases,737 proteases,328 phosphotriesterases,738 and dehydrogenases739 have been engineered using this method. Other methods to detect substrate and product, as well as advances to make the technology more broadly accessible, are being actively developed.740 For example, droplets can be analyzed by using mass spectrometry-based methods, enabling detection of product and substrate without labeling.741 A secondary way to compartmentalize CFE reactions is to use liposomes.620 The main advantage of this method is that it can be used to screen membrane proteins, which may not be folded correctly or active in emulsion-based methods.620 However, this is also compatible with soluble, globular enzymes such as aminoacyl-tRNA synthetases477 and β-glucuronidase.742 These droplet-based technologies are significant advances for enzyme engineering.
Looking forward, cell-free enzyme screening holds great potential for high-throughput homologue comparisons from phylogenetic trees,743 mutagenesis or engineering campaigns,744 and validating de novo design efforts.745 Key application areas include biological synthesis of therapeutic compounds, carbon capture and storage, as well as degradation of environmental contaminants, such as microplastics.744,746
3.8. Enzymatic Cascades and Metabolic Engineering
In addition to screening individual enzymes, cell-free systems provide avenues for multistep biotransformations and biomanufacturing with higher throughput and/or higher efficiency than cellular methods.747,748
3.8.1. Purified Enzymes for Pathways
Purified systems provide efficient routes for biochemical conversion outside the constraints of cells and enable predictable, high-titer processes for industrial chemicals with defined reaction parameters.747−749 These systems can recapitulate pathways from nature,5 or they may use a subset of enzymes for more direct biosynthetic pathways.750 Although purified systems typically use cellular expression rather than CFE, many of the same principles apply for cell-free metabolism in purified and crude systems, as discussed below. Purified enzymes enable the combination of catabolic and anabolic enzymes from diverse clades without optimizing strains or multigene expression cassettes, enabling up to 99% conversion of sugars into 1,4-butanediol or alpha-ketoglutarate751 and the production of cannabinoids at titers 2 orders of magnitude greater than in vivo platforms.752 Other significant products synthesized through the combination of known pathways in vitro include isobutanol at 5 g/L/h,753 monoterpenes at 15 g/L/h with yields greater than 95% through 27 enzymes,754 and myo-inositol produced from starch in 20,000-L reactors.548
Purified enzyme systems also provide the ability to move beyond known metabolic pathways and derive non-natural pathways.755−757De novo pathway design is often bolstered by computational methods that identify enzymes for the desired reactions758,759 and account for the thermodynamics and cofactor requirements of each step, which can increase the theoretical efficiency of synthetic pathways beyond those observed in nature.760 Examples of de novo pathways implemented with purified enzymes include nonoxidative glycolysis to avoid the carbon loss that occurs in natural glycolysis,761 reverse β-oxidation for the synthesis of fatty acids,762,763 the formolase pathway for incorporation of 1-carbon units into metabolism,764 and the CETCH (crotonyl-CoA/ethylmalonyl-CoA/hydroxybutyryl-CoA) cycle for more efficient carbon fixation than the natural Calvin cycle.765 In addition to combining disparate natural catalysts into pathways, purified systems can be interfaced with inorganic catalysts to enable a broader diversity of chemical reactions than what is possible in cells.749 This chemoenzymatic approach is demonstrated by an artificial starch anabolic pathway that employs ZnO-ZrO2 to hydrogenate CO2 to methanol and an enzyme cascade to convert the methanol to starch.764
The primary limitations in purified systems are the cost of enzyme preparation and the need to recycle cofactors and/or byproducts. Purification costs may be reduced through parallelized workflows231 or the use of thermotolerant enzymes such that endogenous proteins can be removed through heat denaturation.548,766 Cofactor recycling in purified systems may simply require additional enzymes,767,768 but orthogonal cofactor specificities can also be implemented.769 Significant examples of in vitro recycling strategies include salvage pathways to recover nonproductive byproducts in the CETCH carbon fixation cycle765 and for polyhydroxybutyrate synthesis770 as well as cofactor pool regulation through a redox purge valve for NAD(P)H771 or an ATP rheostat.772 These strategies bring an element of homeostatic regulation to purified systems.547,747 In addition, one should consider the energy density of substrates and products to enable targeted molecular transformations.
3.8.2. Extract-Based Metabolite Synthesis
While purification offers substantial benefits for enzyme characterization and large-scale biosynthesis, extract-based approaches have utility in prototyping and potential for biomanufacturing. Enzymes in lysate remain active in the soluble fraction of cell extract which was shown to enable conversion of glucose to ethanol,3 avoiding purification while retaining cofactors and other endogenous components that facilitate enzyme activity.747 Crude cell extracts thus enable biochemical conversions through interconnected catabolic and anabolic pathways in the absence of cellular barriers, which have been employed for ethanol production with extracts of C. thermocellum to assess metabolic bottlenecks773 and S. cerevisiae to maximize ethanol titers in vitro.774 Similarly, extracts from Streptomyces spp. have elucidated biochemical mechanisms behind natural products, including fluorinated metabolites,775 anticancer drugs,776 and antibiotics.777−781 Endogenous metabolism in extract can also serve as a precursor to heterologous biochemical synthesis, as seen in the cell-free synthesis of dihydroxyacetone phosphate prior to condensation with butanal to form an unnatural monosaccharide.782 Alternatively, the metabolism of cell extract can be primed or altered through tuning of growth conditions and media for increased flux toward biochemicals of interest.84,783,784
Cell-free metabolism in extracts from engineered strains of E. coli or S. cerevisiae can also produce heterologous metabolites, such as 2,3-butanediol and itaconic acid, with high volumetric productivities and highlight bottlenecks in metabolic flux785,786 while displaying robustness to substrates,787 products,788 temperatures,774 and pH ranges789 that are acutely toxic to living cells.790 This approach also extends to nonmodel organisms, such as the in vitro synthesis of 3-hydroxypropionate in extracts of engineered Pyrococcus furiosus.791 Furthermore, the simplicity of generating extracts from various species enables hybrid cell-free reactions derived from multiple organisms to take advantage of each species’ unique biochemical properties in a manner similar to synthetic microbial consortia without the need to balance interspecies growth.792,793 Although several studies have incorporated mixed extracts for gene expression,56,111,297,794,795 the primary example of hybrid cell-free metabolism combines cyanobacterial extract for starch catabolism with E. coli extract for 2,3-butanediol synthesis.796
The combination of rapid biochemical conversions and increased tolerance to harsher conditions has generated interest in cell extracts as potential biomanufacturing platforms.547,747,797 Although the retention of endogenous metabolism in cell extracts facilitates substrate conversion and cofactor recycling, other endogenous enzymes can drain key intermediates or cofactors from the desired pathway, which limits product yields similarly to cellular systems. Such competing enzymes can be knocked out in the source strain prior to lysis782,798 or (in the case of enzymes essential for growth) selectively removed from the extract through protease cleavage or purification.507,799 Enriched extracts may also be heat treated for thermal “purification” of thermotolerant enzymes,548,766 which represents an intermediary format between fully purified and extract-based cell-free systems. Rather than expressing complete heterologous pathways in cells prior to lysis, many approaches utilize extracts each enriched with one overexpressed enzyme to screen enzyme variants in the context of native metabolism while saving both time and capital costs.800,801 A similar strategy has been employed for laboratory enzymes, such as polymerases and nucleases, to avoid purification.802,803
Mixing extracts individually enriched with a single enzyme has been shown to be a powerful approach for in vitro metabolite synthesis. As compared to using extracts from source cells with multiple enzymes overexepressed, this provides greater control over both the identity and relative concentration of catalyst for each step of the pathway. Extract mixing has been employed to study mevalonate,798 indole-3-acetic acid,804 butanol,731,789 limonene,805 pinene,806 and chlorogenic acid807 synthesis and to explore cell-free metabolism.284 Of note, the number of enzyme variants that can be tested in vitro becomes limited by the generation of many unique cell extracts, the ease of which can vary significantly depending on the scale of available lysis equipment.20
Increased flexibility comes from making a blank slate extract capable of CFE, which can generate libraries of enzyme variants without engineering multiple strains (Figure 13). A complete pathway can be expressed in one reaction using multiple plasmids (relying on relative expression levels similar to in vivo expression),731,808 or each enzyme can be expressed separately and quantified prior to combining the pathway (providing greater precision akin to purified systems).809 Although reagents and metabolic byproducts carried over from protein synthesis can suppress activity of heterologous pathways,284 the ability to rapidly test large combinatorial libraries of enzymes can outweigh reduced titers by significantly decreasing the length of design-build-test-learn cycles. Cell-free prototyping in E. coli extract enabled screening of over 200 pathway variants for 3-hydroxybutryate and butanol biosynthesis to rapidly inform pathway design in C. autoethanogenum production strains.810 Similarly, CFE of 9 heterologous enzymes facilitated the screening of nearly 600 defined pathways for terpene synthesis from glucose to improve limonene titers 25-fold.811 CFE also enables screening enzymes that compete with a pathway of interest. By combining enzymes from C. autoethanogenum suspected to compete with heterologous acetone biosynthesis, an E. coli cell-free system identified targets for genomic knockout in the clostridial production strain that substantially increased selectivity for acetone over undesired byproducts.812 In addition to these cell-free prototyping efforts, the versatility of CFE for metabolite synthesis is highlighted by the wide array of biochemicals synthesized to date. This includes butanol,810 acetone,812 terpenes (limonene, pinene, and bisabolene),811 phenol,84 3-hydroxybutyrate (from glucose810 or whey813), styrene,788 and valinomycin814 with E. coli extracts, unnatural indole alkaloids using PURExpress,815 and heme using S. venezuelae CFE.62
Figure 13.
Generalized workflow for cell-free pathway prototyping. Cell-free reactions incorporating CFE with metabolism simplifies high-throughput screening of enzyme libraries and enables control over many more reaction parameters than in vivo experiments allow.
The diversity of formats and flexibility of applications for cell-free metabolite synthesis presents benefits for high-throughput prototyping and high-yield biotransformations.730,748 These cell-free approaches complement traditional in vivo metabolic engineering and offer the ability to design biosynthesis platforms with increasing speed and precision to facilitate the sustainable production of chemicals at a range of scales.816 Addressing cost and scale-up remain the primary hurdles to transition cell-free biosynthesis from a pathway prototyping method to a biomanufacturing platform.547,747,817
3.9. Building and Engineering Complex Biological Systems
Cell-free systems derive benefits from the lack of viability constraints, but defining features of cells, such as compartmentalization and self-replication, can also be valuable in nonliving contexts. Research into artificial cells spans many disciplines and applications, generally focusing on top-down reduction of cells into nonliving protocells or bottom-up generation of complex systems.29,818,819 Although the abiotic context provides immense control over the composition,820 functionalization,821 and contents822 of artificial cells, the goal of self-replication remains difficult to achieve in full.823 Here we briefly discuss bottom-up efforts to build cell-like structures and complex assemblies, including ribosomes and viruses.
3.9.1. Building Ribosomes
Understanding and engineering ribosomes could repurpose protein translation for synthetic biology applications. The prokaryotic ribosome is a complex molecular machine comprising three ribosomal RNAs (rRNAs) and 54 ribosomal proteins. These assemble into a 70S particle from two parts: the 30S small subunit, comprising the 16S rRNA and 21 r-proteins, and the 50S large subunit, comprising the 23S rRNA, 33 r-proteins, and the 5S rRNA. The total set of r-proteins is referred to as TP70. Ribosome assembly consists of transcription of the rRNA as a single transcript, cleavage of the rRNA transcript to yield the 23S, 16S, and 5S rRNA, maturation of the rRNA involving additional processing and post-transcriptional modifications, and then assembly of individual subunits through well-studied assembly pathways with the r-proteins.824 Below we describe efforts in ribosome assembly and synthesis, with an emphasis on E. coli ribosomes.
An early example of small subunit reconstitution was published in 1968 using 16S rRNA and r-proteins that were isolated from purified small subunits.825 Recombinant r-proteins were sufficient to reconstitute small subunit assembly when mixed with native 16S rRNA.826 Similarly, in vitro transcribed 16S rRNA could be mixed with native r-proteins to form active small subunits.827,828 Later experiments showed that the small subunit could be assembled using components purified from nonribosomal sources.829 Assembly can also be assisted by the addition of ribosome biogenesis factors or chaperones that facilitate folding and assembly at lower temperatures.827,830,831 In total, this shows that the small subunit can be assembled into functional particles within in vitro environments.
Reconstitution of the large subunit has proven to be more difficult than that of the small subunit, likely due to the size of the 23S rRNA, essential post-transcriptional modifications of the rRNA, and the complexity of its interactions with the 33 r-proteins. Like the 30S small subunit, reconstitution of the large subunit was first successfully achieved with 23S rRNA, 5S rRNA, and r-proteins that were purified from native 50S subunits.832,833 R-proteins from recombinant sources could also be used to reconstitute 50S ribosomes with 23S and 5S rRNA isolated from 50S subunits.834 However, 50S subunits could only be reassembled with minimal activity using in vitro transcribed 23S rRNA.835 This assembly defect is hypothesized to arise from the lack of post-transcriptional modifications on the rRNA. Interestingly, small molecules such as osmolytes and antibiotics were found to improve assembly of 50S subunits from in vitro transcribed 23S rRNAs.836
Methods to test ribosome assembly and function have also been developed. One-pot integrated rRNA Synthesis, Assembly, and Translation (iSAT; Figure 14)837 consists of a cellular extract that is depleted of wild-type ribosomes through ultracentrifugation, a plasmid encoding an rRNA operon under the control of a T7 promoter, a plasmid containing an appropriate reporter such as luciferase or sfGFP, and TP70 purified from native ribosomes. All steps of iSAT take place at 37 °C, physiological pH, in buffer and salt conditions that mimic the E. coli cytoplasm. Furthermore, the reaction formulation can be changed to enable a wide variety of applications. iSAT enables ribosomes to be synthesized and tested for function, providing a workflow for engineering ribosomes.
Figure 14.

Ribosome synthesis and assembly using the iSAT system.
A variety of advances have enabled iSAT to become a powerful technology for ribosome synthesis, assembly, and function. For instance, running iSAT reactions in fed-batch mode or in semicontinuous mode increased reaction duration and protein yields by providing necessary phosphoenolpyruvate and NTPs.838 Additional optimizations included modulating rRNA transcription by adding 3′ modifications to the rRNA transcript and combinatorally optimizing DNA concentrations.839 Crowding agents and reducing conditions were also found to improve iSAT yields.840 In total, these optimizations have enabled sfGFP yields at ∼8 μM, or ∼220 μg/mL. iSAT has been modified to run in PURE-like conditions to engineer the 30S subunit, as post-transcriptional modifications of the 16S rRNA are not required for assembly.828,829 In addition, iSAT enables the study and engineering of ribosomes that are competent for translation but unable to support life in laboratory conditions.289,841 For example, toward characterizing the mutational flexibility of the active site of the ribosome, the peptidyl transferase center (PTC), ribosomal variants with mutations throughout the PTC were constructed and characterized in iSAT.841,842 Further, iSAT can be coupled to screening technologies, such as ribosome display and in vitro compartmentalization, to evolve the ribosome to enable improved function (e.g., resistance toward antibiotics, improved assembly of active subunits).289,828,843 iSAT has also been used to advance the concept of “ribosome pool engineering” by identifying sequence-optimized ribosomes with improved protein biosynthesis yields.844 These advances show how cell-free systems can be used to study and engineer ribosomes for basic science and engineering.
3.9.2. Reconstituting Bacteriophages
Cell-free synthesis of viruses represents a simpler synthetic system than artificial cells due to self-assembly and the lack of membrane chemistry requirements. In addition to enabling basic research of viral assembly without passaging or infecting cells, CFE could serve as a production platform for virus-like particles (VLPs) or phage therapies to target cancer, antibiotic resistance, and other challenges in human health and agriculture.172 Self-assembling viral capsids have been produced in eukaryotic cell extracts since the 1980s, either through translation of in vitro transcribed RNA or through coupled transcription and translation from DNA. For example, hepatitis B and C capsids were expressed in extracts from rabbit reticulocyte,845 wheat germ,846 and P. pastoris(847) to study assembly and molecular regulation. Similar studies were carried out for cell-free synthesis and assembly of herpes simplex virus capsids in S. frugiperda extracts848,849 and vesicular stomatitis virus in rabbit reticulocyte extracts.850 In 2020, CFE with extract from L. tarentolae provided a platform to study protein interactions important for the assembly of Zika virus.851 In addition to these eukaryotic examples, E. coli extracts have synthesized MS2 and hepatitis B capsids up to 1013 particles/mL.852
Beyond VLPs and capsids, cell-free systems have demonstrated the capacity to replicate and/or synthesize complete viruses from eukaryotes and prokaryotes. Although the in vitro replication of tomato bushy stunt virus in S. cerevisiae extract was dependent on membrane vesicles,853 the cell-free synthesis of infectious viruses indicates that living cells with large segments of membrane are not required for viral reproduction.172 The best explored virus classes in CFE are eukaryote-infecting picornaviruses and E. coli bacteriophages, which demonstrate the ability of cell-free systems to carry out large genetic programs and build biological machines. Poliovirus, with a 7.5 kb RNA genome, was the first virus fully reconstituted in an extract-based cell-free system,854 with expression in HeLa cell extracts achieving titers up to 107 infectious particles/mL.855−857 Similarly, the synthesis of encephalomyocarditis virus, with a 7.8 kb RNA genome, was demonstrated in several eukaryotic cell extracts with titers up to 107 particles/mL.858−860
Optimized E. coli CFE systems have synthesized high titers of increasingly complex phages up to 1013 particles/mL.21 Successfully expressed and assembled phages include MS2 (3.6 kb RNA genome), ΦX174 (5.6 kb ssDNA genome), T7 (40 kb dsDNA genome), and T4 (170 kb dsDNA genome).44,172,861,862 Cell-free phage synthesis has also been employed as a quality control method to compare the activity of cell extracts prepared with different methods86 and applied as a screening and engineering platform for potential therapeutics.863−865 Although this application space has only explored a small fraction of viral diversity to date, cell-free systems offer the potential to synthesize the vast number of sequenced viruses (from both prokaryotes866 and eukaryotes867) for basic and applied research without precise knowledge of host ranges or the ability to culture host organisms. Researchers, however, must remain mindful of ethical and biosafety considerations around viral synthesis, especially for eukaryotic viruses that can infect humans, employing appropriate safeguards and including critical discussion in publications.868 In addition, a key caution is to ensure that extracts used for phage assembly are free from any living cells.
3.9.3. Building Synthetic Cells
This section focuses on key aspects of synthetic cell research incorporating CFE systems; other aspects of this broad field have been recently reviewed elsewhere.869 The compartmentalization of CFE systems in droplets or vesicles begins to transition these systems toward cell-like structures capable of replicating genetic information and carrying out the steps of the central dogma823 while physically linking genotype to phenotype.323 The functionality of these synthetic cells is determined by the encapsulation mechanism (including polymers, oils, and liposomes; see Section 2.5), the presence and type of protein or nucleic acid functionalization, and the interior composition (Figure 15). Artificial cells based on CFE systems have facilitated a wide range of outputs from continuous reporter protein expression47 to cytotoxic protein synthesis within mouse models870 and the parallelized screening of antimicrobial peptides871 and ncAAs.872 Furthermore, CFE assembled in compartments on microfluidic chips can provide a platform for synthetic cell studies with movement-restricted reaction arrays. Demonstrations using cell extracts diffusing into compartments coated with DNA have shown cell-like protein expression gradients,245 temporal propagation of expression,314 and the ability to tune logic gate speed or precision.313
Figure 15.
Variables under control for artificial cells. Designing functionalized liposomes enables CFE reactions to more closely resemble cells with complex, context-dependent behaviors. Common variables are highlighted in orange.
Encapsulated reactions more closely resemble cells when they generate proteins that interact with the surrounding membrane or execute behaviors such as communication and differentiation. Several cytoskeletal proteins have been synthesized within artificial cells to alter liposome shape or stability, including MreB,873,874 tubulins,875 and actin.876 Additionally, the expression of cell division proteins FtsZ, FtsA, and ZipA within liposomes causes deformation or constriction that could be incorporated along with assembly regulating proteins877 into a minimal division mechanism.878−880 In contrast, communication between artificial cells is a more explored research area due to the rich diversity of approaches using functionalized membranes, protein–protein interactions, and signaling pathways to facilitate programs within881 and between encapsulated cell-free systems.882,883 Inducible promoters676 and expression cascades247 enable context-dependent execution of genetic programs within liposomes, often with the assistance of α-hemolysin for selective permeability or as a component of logic gates.884,885 The fusion of distinct artificial cell populations can also be mediated by SNARE proteins886 or complementary DNA strands on the exterior of liposomes to facilitate more complex or interaction-dependent behaviors,887 such as genetic or metabolic differentiation.885,886 Furthermore, biochemical communication can extend between living and artificial cells, since the mechanisms implemented in liposomes originally evolved in nature.888 This typically consists of the exchange of quorum sensing molecules between artificial and living cells in the same solution889−892 or the implementation of signaling cascades that trigger cellular responses, such as protein synthesis893 or differentiation.894 Alternatively, the construction of large liposomes enables the encapsulation of whole cells and other components to form “hybrid cells” capable of layered functionalities.888,895
In addition to intercellular communication, the ability to sense and respond to environmental stimuli is a hallmark of life. With this feature in mind, artificial cells have been engineered to respond to changes in pH, osmotic pressure, and light. For example, acid-induced phase separation increased the activity of encapsulated enzymes by raising the local concentration of both enzyme and substrate,896 and metal–organic frameworks facilitated pH-gated gene expression from encapsulated HeLa cell extract.897 Pressure-responsive artificial cells have incorporated the membrane protein MscL (large-conductance mechanosensitive channel) coupled with nested vesicles898 or biosensor cascades899,900 to link biochemical programs to changes in osmotic pressure, membrane asymmetry,901 and chemical inducers. This channel has also facilitated studies of membrane protein incorporation into vesicles by CFE.394,902 Additionally, the functionalization of synthetic cell membrane or contents enables light-driven reactions without photoautotrophic cells. These systems typically employ bacteriorhodopsin to generate light-induced proton gradients coupled with ATPase for ATP generation,903 enabling energy-intensive processes including actin polymerization876 and combined transcription-translation.904 Alternatively, artificial cells may incorporate photosynthetic organelles to for light-powered gene expression905 or enzyme cascades906 through the combination of natural and synthetic parts.
Despite the ability for compartmentalized CFE reactions to produce a wide range of proteins and execute complex programs, self-replication remains a challenging goal for artificial cell systems.823 DNA replication in vesicles has been demonstrated successfully alongside protein synthesis,907,908 but resource limitations in PURE- and extract-based CFE systems limit the extent to which the full complement of gene expression machinery can be regenerated simultaneously.261,355,909 Optimization of gene expression in artificial cells using highly tuned CFE recipes and/or permeable liposomes in a feeding solution for semicontinuous gene expression show promise,21 but the dilute environment of cell-free reactions relative to cytoplasm results in lower protein synthesis rates than observed in living cells.19,179 Serial transfer of PURE reactions enabled the reconstitution of RNA polymerase activity and several key enzymes for translation,910 and simultaneous generation of RNA polymerase and aminoacyl-tRNA synthetases was enabled by extensively tuning template concentrations.911 Yet, a major bottleneck remains in regenerating all components with high activity912 and synthesizing ribosomes fully in vitro. Protein synthesis has been demonstrated in liposomes containing ribosomes assembled in vitro after harvesting the components from cells,843 but ribosomal proteins synthesized by PURE show reduced activity,913 potentially due to the lack of certain assembly cofactors and/or rRNA modifications. However, promising steps have been made toward synthesizing liposomes with both soluble914 and membrane-bound412,915 enzymes that could form the basis of replicating artificial cells. Looking ahead, there is a need to integrate artificial cells with functional cell division mechanisms that operate in concert with engineered reactions within the mother and daughter particles. Understanding and engineering smaller self-assembling biological systems, such as viruses and ribosomes, will facilitate further breakthroughs in artificial cell optimization toward fully functional self-replicating systems.
3.10. Biology Education
In addition to numerous opportunities in basic and applied research, cell-free synthetic biology provides a unique platform for education in classrooms ranging from middle school to undergraduate levels. Enzymes provide discrete examples of laboratory-based education for biochemistry, ranging from simple assays to more complex analyses of reaction kinetics.916,917 However, more complex concepts in molecular and synthetic biology are traditionally taught through lectures alone due to the cost and complexity of cell-based laboratories. This limits hands-on demonstrations and research experiences to well-funded high schools and colleges, typically in extracurricular groups such as the international Genetically Engineered Machines (iGEM) competition918 and BioBuilder.919 Cell-free approaches have led to the development of hands-on education modules with the potential for widespread implementation of engaging synthetic biology education due to the minimal equipment and biosafety protocols required.920 Extract-based approaches with combined transcription-translation enable rapid demonstrations of biology’s central dogma through visual reporters, such as fluorescent proteins, that can be seen through inexpensive imagers or the naked eye.145 More advanced educational kits have explored optogenetics,921 CRISPR-Cas9 and antibiotic resistance,922−924 CRISPRi,925 environmental sampling,667 or multiple sensory outputs (visual, olfactory, and tactile)144 to highlight the versatility of biological systems. Furthermore, lyophilization of cell-free reactions increases shelf stability for at least 6 months in a refrigerator, which provides the potential for widespread distribution of low-cost biology education kits around the world from a centralized production facility prior to activation through the simple addition of water and/or DNA. As cell-free biology continues to advance, these systems are poised to transform synthetic biology education by facilitating earlier and more widespread exposure to the field and the rapidly expanding biotechnology industry.926−928
4. Perspective
CFE systems have significantly advanced over the past 20 years, facilitating numerous applications. As we look to the future, we anticipate several key trends that will shape the evolution of the CFE field.
Increasing CFE reaction yields, reaction longevity, and the diversity of functionally expressible proteins will open completely new use-inspired applications. A CFE batch reaction yield of >10 mg/mL, which is sufficient for self-replication of E. coli,30 will provide new opportunities for large-scale manufacturing and enable applications that require coexpression of multiple genes, such as prototyping molecular assemblies, metabolic networks, and synthetic cells. Achieving low-cost and long-lived reactions will require research into why reactions stop (e.g., via omics methods81,84−90), exogenous energy sources (e.g., light904 and electricity929), and alternative organisms. Beyond higher production yields, a key goal should be the manufacture of any proteoform, encompassing all sequence and splice variants of a protein and its numerous possible post-translational modifications.930 While efforts in glycosylation have advanced, many post-translational modifications remain underrepresented. Indeed, we expect the community to push toward the creation of a universal in vitro protein manufacturing system, enabling researchers to choose their preferred translation rates, folding environments, post-translational modifications, and other relevant parameters. The capability to quickly produce any defined proteoform would transform efforts to comprehend and engineer protein function.
Although adoption of CFE systems is increasing, their use in most laboratories is far from routine. To make CFE systems ubiquitous, the community must improve protocol clarity and streamline reagent preparation. Reproducible reagent and cell-extract production across laboratories can be a challenge,127,210 and most current protocols involve the preparation of cell extract and dozens of reagents, which can be daunting for new researchers. In the past five years, progress has been made in lowering the barrier to entry, with new user-focused primers,30 video-based protocols,931 simplified E. coli cell-extract preparation protocols,87 and efforts to understand the sources of variability in CFE system manufacture.208,724 However, more work remains to create repeatable, streamlined protocols and CFE systems for routine use outside of expert CFE laboratories. The availability of low cost (i.e., rivaling cost-effective in-house systems), high-yielding commercial CFE systems, would also increase usage and accessibility.
CFE systems are well positioned to generate high-quality data to support the development of biophysical or artificial intelligence models capable of predicting biological function. Through integration with automation, microfabrication, microfluidics, or display techniques, CFE can enable the rapid and regular evaluation of thousands to millions of unique conditions.319,320,329,344,604
CFE-based biosensors hold potential as field-deployable and on-demand diagnostic platforms. These sensors will need to demonstrate robust detection of analytes with relevant limits of detection and signal activation outside of the laboratory setting. Identification of new biological sensors, both new analytes that can be measured as well as new sensing modalities, will expand use cases for CFE sensors. Compatibility with complex sample matrices (e.g., wastewater, blood) will be crucial for detecting human health biomarkers and environmental contaminants.
We expect that new distributed manufacturing opportunities will emerge and lead to an entirely new biotechnology industry. Scale-up of protein and biomolecule manufacturing remains relatively underexplored. A key roadblock has been high costs and bioprocess engineering know-how. We anticipate that innovations to further reduce costs as well as purification and downstream quality control processes will increase scale-up applications and distribution of cell-free expression reactions. The development of new materials such as paper, fabric, and hydrogels for distribution is also poised for impact in point-of-use settings.
CFE systems will remain essential tools for studying and building biological systems. Like their early use, CFE systems will continue to contribute to the study and expansion of the genetic code, particularly through the incorporation of noncanonical amino acids into proteins. Efforts to build synthetic cells represent another exciting frontier, with many core functions demonstrated in isolated environments. Future efforts will focus on integrating these functions together and improving their efficiencies to realize fully functional synthetic cells.
The chapters above highlight the enormous growth of CFE systems and their applications. In the coming decade, CFE systems will continue enabling researchers to explore biological systems beyond the cell, transform synthetic biology, and accelerate scientific discovery to address pressing societal challenges.
Acknowledgments
We would like to thank the reviewers for providing invaluable feedback, which improved the accuracy and quality of this work. It was a nontrivial task to review a manuscript of this size and they were both constructive and rigorous in their work. We also thank all members of the lab, past and present, and the Northwestern Center for Synthetic Biology for invaluable discussions that have contributed to the ideas of this paper. This work is supported by the Army Research Laboratory and the Army Research Office (W911NF-23-1-0334, W911NF-22-2-0210, W911NF-22-2-0246, W911NF-18-1-0200), Army Contracting Command (W52P1J-21-9-3023), DTRA (HDTRA1-21-1-0038, HDTRA1-20-1-0004) the Department of Energy (DE-SC0023278, DE-NA0003525), the National Science Foundation (CBET - 1936789), the Gates Foundation (INV-038694), DARPA (W911NF-23-2-0039), and the National Institutes of Health (1U19AI142780-01).
Biographies
Andrew Hunt received a B.S. in Chemical Engineering from The New Mexico Institute of Mining and Technology and a Ph.D. in Chemical and Biological Engineering from Northwestern University under the guidance of Dr. Michael Jewett. His Ph.D. research focused on the development of high-throughput screening platforms for protein–protein interactions leveraging cell-free gene expression systems. He is currently a postdoctoral fellow at the University of Washington, where he is using computational protein design to create new enzymes.
Blake Rasor earned a B.S. in Biology and Microbiology from Miami University and a Ph.D. in Chemical and Biological Engineering from Northwestern University, where his research focused on metabolism and pathway prototyping in cell-free systems. He is currently a postdoctoral fellow at the Max Planck Institute for Terrestrial Microbiology applying cell-free methodologies to study and engineer photosynthetic systems.
Kosuke Seki is currently a postdoctoral fellow at the University of California, San Francisco under Dr. Tanja Kortemme. He received a Ph.D in Chemical and Biological Engineering with Dr. Michael Jewett using cell-free gene expression systems for genetic code expansion, and a B.S. in Chemical Engineering at the University of California, Irvine. He is currently interested in studying structure–function relationships in proteins.
Holly Ekas received a B.S. in Chemical Engineering from The University of Texas at Austin and a Ph.D. in Chemical and Biological Engineering from Northwestern University under Dr. Michael Jewett. Her Ph.D. work focused on developing high-throughput engineering platforms for cell-free biosensor development.
Katherine Warfel received a B.S. in Chemical Engineering from Bucknell University and a Ph.D in Chemical and Biological Engineering from Northwestern University under the guidance of Dr. Michael Jewett. Her Ph.D. research leveraged cell-free gene expression systems to synthesize glycoconjugate vaccines. Through this work, she engineered membrane-bound glycosylation systems, and optimized the cost and thermostability of cell-free reactions to enhance the production of glycosylated protein vaccines.
Ashty Karim is an assistant professor of Chemical and Biological Engineering at Northwestern University. He earned his B.S. degrees in Chemical Engineering and in Biology from the University of Texas at Austin and received his Ph.D. in Chemical Engineering from Northwestern University where he developed cell-free systems to build and prototype enzymes and enzymatic cascades. Ashty works at the interface of biology and chemistry to develop synthetic biology solutions in sustainability.
Michael Jewett is a Professor of Bioengineering at Stanford University. His research group focuses on advancing synthetic biology research to support planet and societal health, with applications in medicine, manufacturing, sustainability, and education. Dr. Jewett received his PhD in 2005 at Stanford University and completed postdoctoral studies at the Center for Microbial Biotechnology in Denmark and the Harvard Medical School. He began his academic career at Northwestern University in 2009, where he was the Director of the Center for Synthetic Biology.
Author Contributions
¶ A.C.H. and B.J.R. contributed equally.
The authors declare the following competing financial interest(s): M.C.J. is a co-founder and has financial interest in Stemloop, Inc., Pearl Bio, Gauntlet Bio, and Synolo Therapeutics. These interests are reviewed and managed by Northwestern University and Stanford University in accordance with their conflict of interest policies.
References
- Silverman A. D.; Karim A. S.; Jewett M. C. Cell-Free Gene Expression: An Expanded Repertoire of Applications. Nat. Rev. Genet. 2020, 21 (3), 151–170. 10.1038/s41576-019-0186-3. [DOI] [PubMed] [Google Scholar]
- Chong S. Overview of Cell-Free Protein Synthesis: Historic Landmarks, Commercial Systems, and Expanding Applications. Curr. Protoc. Mol. Biol. 2014, 108 (1), 16.30.1–11. 10.1002/0471142727.mb1630s108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Buchner E. Alkoholische Gährung Ohne Hefezellen. Ber. Dtsch. Chem. Ges. 1897, 30 (1), 117–124. 10.1002/cber.18970300121. [DOI] [Google Scholar]
- Kohler R. E. The Reception of Eduard Buchner’s Discovery of Cell-Free Fermentation. J. Hist. Biol. 1972, 5 (2), 327–353. 10.1007/BF00346663. [DOI] [PubMed] [Google Scholar]
- Welch P.; Scopes R. K. Studies on Cell-Free Metabolism: Ethanol Production by a Yeast Glycolytic System Reconstituted from Purified Enzymes. J. Biotechnol. 1985, 2 (5), 257–273. 10.1016/0168-1656(85)90029-X. [DOI] [Google Scholar]
- Matthews G.; Colman A. A Highly Efficient, Cell-Free Translation/Translocation System Prepared from Xenopus Eggs. Nucleic Acids Res. 1991, 19 (23), 6405–6412. 10.1093/nar/19.23.6405. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nirenberg M. W.; Matthaei J. H. The Dependence of Cell-Free Protein Synthesis in E. Coli upon Naturally Occurring or Synthetic Polyribonucleotides. Proc. Natl. Acad. Sci. U. S. A. 1961, 47 (10), 1588–1602. 10.1073/pnas.47.10.1588. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nirenberg M.; Leder P.; Bernfield M.; Brimacombe R.; Trupin J.; Rottman F.; O’Neal C. RNA Codewords and Protein Synthesis, VII. On the General Nature of the RNA Code. Proc. Natl. Acad. Sci. U. S. A. 1965, 53 (5), 1161–1168. 10.1073/pnas.53.5.1161. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zamecnik P. C.; Frantz I. D.; Loftfield R. B.; Stephenson M. L. Incorporation in Vitro of Radioactive Carbon from Carboxyl-Labeled Dl-Alanine and Glycine into Proteins of Normal and Malignant Rat Livers. J. Biol. Chem. 1948, 175 (1), 299–314. 10.1016/S0021-9258(18)57260-4. [DOI] [PubMed] [Google Scholar]
- Littlefield J. W.; Keller E. B.; Gross J.; Zamecnik P. C. Studies on Cytoplasmic Ribonucleoprotein Particles from the Liver of the Rat. J. Biol. Chem. 1955, 217 (1), 111–123. 10.1016/S0021-9258(19)57162-9. [DOI] [PubMed] [Google Scholar]
- Hoagland M. B.; Keller E. B.; Zamecnik P. C. Enzymatic Carboxyl Activation of Amino Acids. J. Biol. Chem. 1956, 218 (1), 345–358. 10.1016/S0021-9258(18)65898-3. [DOI] [PubMed] [Google Scholar]
- Gale E. F.; Folkes J. P. Effect of Nucleic Acids on Protein Synthesis and Amino-Acid Incorporation in Disrupted Staphylococcal Cells. Nature 1954, 173 (4417), 1223–1227. 10.1038/1731223a0. [DOI] [PubMed] [Google Scholar]
- Caskey C. T.; Tompkins R.; Scolnick E.; Caryk T.; Nirenberg M. Sequential Translation of Trinucleotide Codons for the Initiation and Termination of Protein Synthesis. Science 1968, 162 (3849), 135–138. 10.1126/science.162.3849.135. [DOI] [PubMed] [Google Scholar]
- Zubay G. In Vitro Synthesis of Protein in Microbial Systems. Annu. Rev. Genet. 1973, 7 (1), 267–287. 10.1146/annurev.ge.07.120173.001411. [DOI] [PubMed] [Google Scholar]
- Nevin D. E.; Pratt J. M. A Coupled in Vitro Transcription-Translation System for the Exclusive Synthesis of Polypeptides Expressed from the T7 Promoter. FEBS Lett. 1991, 291 (2), 259–263. 10.1016/0014-5793(91)81297-L. [DOI] [PubMed] [Google Scholar]
- Krieg P. A.; Melton D. A. In Vitro RNA Synthesis with SP6 RNA Polymerase. Methods Enzymol. 1987, 155, 397–415. 10.1016/0076-6879(87)55027-3. [DOI] [PubMed] [Google Scholar]
- Kim D. M.; Choi C. Y. A Semicontinuous Prokaryotic Coupled Transcription/Translation System Using a Dialysis Membrane. Biotechnol. Prog. 1996, 12 (5), 645–649. 10.1021/bp960052l. [DOI] [PubMed] [Google Scholar]
- Jewett M. C.; Swartz J. R. Mimicking the Escherichia Coli Cytoplasmic Environment Activates Long-Lived and Efficient Cell-Free Protein Synthesis. Biotechnol. Bioeng. 2004, 86 (1), 19–26. 10.1002/bit.20026. [DOI] [PubMed] [Google Scholar]
- Jewett M. C.; Calhoun K. A.; Voloshin A.; Wuu J. J.; Swartz J. R. An Integrated Cell-Free Metabolic Platform for Protein Production and Synthetic Biology. Mol. Syst. Biol. 2008, 4 (1), 220. 10.1038/msb.2008.57. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cole S. D.; Miklos A. E.; Chiao A. C.; Sun Z. Z.; Lux M. W. Methodologies for Preparation of Prokaryotic Extracts for Cell-Free Expression Systems 2020, 5, 252–267. 10.1016/j.synbio.2020.07.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Garenne D.; Thompson S.; Brisson A.; Khakimzhan A.; Noireaux V. The All-E. ColiTXTL Toolbox 3.0: New Capabilities of a Cell-Free Synthetic Biology Platform. Synth. Biol. 2021, 6 (1), ysab017. 10.1093/synbio/ysab017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martin R. W.; Des Soye B. J.; Kwon Y.-C.; Kay J.; Davis R. G.; Thomas P. M.; Majewska N. I.; Chen C. X.; Marcum R. D.; Weiss M. G.; Stoddart A. E.; Amiram M.; Ranji Charna A. K.; Patel J. R.; Isaacs F. J.; Kelleher N. L.; Hong S. H.; Jewett M. C. Cell-Free Protein Synthesis from Genomically Recoded Bacteria Enables Multisite Incorporation of Noncanonical Amino Acids. Nature Communications 2018, 9 (1), 1203. 10.1038/s41467-018-03469-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Des Soye B. J.; Gerbasi V. R.; Thomas P. M.; Kelleher N. L.; Jewett M. C. A Highly Productive, One-Pot Cell-Free Protein Synthesis Platform Based on Genomically Recoded Escherichia Coli. Cell Chem. Biol. 2019, 26 (12), 1743–1754. 10.1016/j.chembiol.2019.10.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Caschera F.; Noireaux V. Synthesis of 2.3 Mg/Ml of Protein with an All Escherichia Coli Cell-Free Transcription-Translation System. Biochimie 2014, 99, 162–168. 10.1016/j.biochi.2013.11.025. [DOI] [PubMed] [Google Scholar]
- Madin K.; Sawasaki T.; Ogasawara T.; Endo Y. A Highly Efficient and Robust Cell-Free Protein Synthesis System Prepared from Wheat Embryos: Plants Apparently Contain a Suicide System Directed at Ribosomes. Proc. Natl. Acad. Sci. U. S. A. 2000, 97 (2), 559–564. 10.1073/pnas.97.2.559. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shimizu Y.; Inoue A.; Tomari Y.; Suzuki T.; Yokogawa T.; Nishikawa K.; Ueda T. Cell-Free Translation Reconstituted with Purified Components. Nat. Biotechnol. 2001, 19 (8), 751–755. 10.1038/90802. [DOI] [PubMed] [Google Scholar]
- Carlson E. D.; Gan R.; Hodgman C. E.; Jewett M. C. Cell-Free Protein Synthesis: Applications Come of Age. Biotechnol. Adv. 2012, 30 (5), 1185–1194. 10.1016/j.biotechadv.2011.09.016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gregorio N. E.; Levine M. Z.; Oza J. P. A User’s Guide to Cell-Free Protein Synthesis. Methods Protoc. 2019, 2 (1), 24. 10.3390/mps2010024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Laohakunakorn N.; Grasemann L.; Lavickova B.; Michielin G.; Shahein A.; Swank Z.; Maerkl S. J. Bottom-up Construction of Complex Biomolecular Systems with Cell-Free Synthetic Biology. Front. Bioeng. Biotechnol. 2020, 8, 213. 10.3389/fbioe.2020.00213. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Garenne D.; Haines M. C.; Romantseva E. F.; Freemont P.; Strychalski E. A.; Noireaux V.. Cell-Free Gene Expression. Nat. Rev. Methods Primers 2021, 1 ( (1), ). 10.1038/s43586-021-00046-x. [DOI] [Google Scholar]
- Noireaux V.; Liu A. P. The New Age of Cell-Free Biology. Annu. Rev. Biomed. Eng. 2020, 22 (1), 51–77. 10.1146/annurev-bioeng-092019-111110. [DOI] [PubMed] [Google Scholar]
- Boles K. S.; Kannan K.; Gill J.; Felderman M.; Gouvis H.; Hubby B.; Kamrud K. I.; Venter J. C.; Gibson D. G. Digital-to-Biological Converter for on-Demand Production of Biologics. Nat. Biotechnol. 2017, 35 (7), 672–675. 10.1038/nbt.3859. [DOI] [PubMed] [Google Scholar]
- Dopp J. L.; Tamiev D. D.; Reuel N. F. Cell-Free Supplement Mixtures: Elucidating the History and Biochemical Utility of Additives Used to Support in Vitro Protein Synthesis in E. Coli Extract. Biotechnol. Adv. 2019, 37 (1), 246–258. 10.1016/j.biotechadv.2018.12.006. [DOI] [PubMed] [Google Scholar]
- Sawasaki T.; Ogasawara T.; Morishita R.; Endo Y. A Cell-Free Protein Synthesis System for High-Throughput Proteomics. Proc. Natl. Acad. Sci. U. S. A. 2002, 99 (23), 14652–14657. 10.1073/pnas.232580399. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Karig D. K.; Jung S.-Y.; Srijanto B.; Collier C. P.; Simpson M. L. Probing Cell-Free Gene Expression Noise in Femtoliter Volumes. ACS Synth. Biol. 2013, 2 (9), 497–505. 10.1021/sb400028c. [DOI] [PubMed] [Google Scholar]
- Shojaeian M.; Lehr F.-X.; Göringer H. U.; Hardt S. On-Demand Production of Femtoliter Drops in Microchannels and Their Use as Biological Reaction Compartments. Anal. Chem. 2019, 91 (5), 3484–3491. 10.1021/acs.analchem.8b05063. [DOI] [PubMed] [Google Scholar]
- Zawada J. F.; Yin G.; Steiner A. R.; Yang J.; Naresh A.; Roy S. M.; Gold D. S.; Heinsohn H. G.; Murray C. J. Microscale to Manufacturing Scale-up of Cell-Free Cytokine Production-a New Approach for Shortening Protein Production Development Timelines. Biotechnol. Bioeng. 2011, 108 (7), 1570–1578. 10.1002/bit.23103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim D. M.; Kigawa T.; Choi C. Y.; Yokoyama S. A Highly Efficient Cell-Free Protein Synthesis System from Escherichia Coli. Eur. J. Biochem. 1996, 239 (3), 881–886. 10.1111/j.1432-1033.1996.0881u.x. [DOI] [PubMed] [Google Scholar]
- Voloshin A. M.; Swartz J. R. Efficient and Scalable Method for Scaling up Cell Free Protein Synthesis in Batch Mode. Biotechnol. Bioeng. 2005, 91 (4), 516–521. 10.1002/bit.20528. [DOI] [PubMed] [Google Scholar]
- Calhoun K. A.; Swartz J. R. Energizing Cell-Free Protein Synthesis with Glucose Metabolism. Biotechnol. Bioeng. 2005, 90 (5), 606–613. 10.1002/bit.20449. [DOI] [PubMed] [Google Scholar]
- Kim H.-C.; Kim T.-W.; Kim D.-M. Prolonged Production of Proteins in a Cell-Free Protein Synthesis System Using Polymeric Carbohydrates as an Energy Source. Process Biochem. 2011, 46 (6), 1366–1369. 10.1016/j.procbio.2011.03.008. [DOI] [Google Scholar]
- Shin J.; Noireaux V. An E. Coli Cell-Free Expression Toolbox: Application to Synthetic Gene Circuits and Artificial Cells. ACS Synth. Biol. 2012, 1 (1), 29–41. 10.1021/sb200016s. [DOI] [PubMed] [Google Scholar]
- Cai Q.; Hanson J. A.; Steiner A. R.; Tran C.; Masikat M. R.; Chen R.; Zawada J. F.; Sato A. K.; Hallam T. J.; Yin G. A Simplified and Robust Protocol for Immunoglobulin Expression in E Scherichia Coli Cell-Free Protein Synthesis Systems. Biotechnol. Prog. 2015, 31 (3), 823–831. 10.1002/btpr.2082. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Garamella J.; Marshall R.; Rustad M.; Noireaux V. The All E. Coli TX-TL Toolbox 2.0: A Platform for Cell-Free Synthetic Biology. ACS Synth. Biol. 2016, 5 (4), 344–355. 10.1021/acssynbio.5b00296. [DOI] [PubMed] [Google Scholar]
- Warfel K. F.; Williams A.; Wong D. A.; Sobol S. E.; Desai P.; Li J.; Chang Y.-F.; DeLisa M. P.; Karim A. S.; Jewett M. C. A Low-Cost, Thermostable, Cell-Free Protein Synthesis Platform for On-Demand Production of Conjugate Vaccines. ACS Synth. Biol. 2023, 12, 95. 10.1021/acssynbio.2c00392. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 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 (1), 15–19. 10.1016/S0014-5793(98)01620-2. [DOI] [PubMed] [Google Scholar]
- Noireaux V.; Libchaber A. A Vesicle Bioreactor as a Step toward an Artificial Cell Assembly. Proc. Natl. Acad. Sci. U. S. A. 2004, 101 (51), 17669–17674. 10.1073/pnas.0408236101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Klammt C.; Schwarz D.; Fendler K.; Haase W.; Dötsch V.; Bernhard F. Evaluation of Detergents for the Soluble Expression of Alpha-Helical and Beta-Barrel-Type Integral Membrane Proteins by a Preparative Scale Individual Cell-Free Expression System. FEBS J. 2005, 272 (23), 6024–6038. 10.1111/j.1742-4658.2005.05002.x. [DOI] [PubMed] [Google Scholar]
- Jackson R. J.; Hunt T.. [4] Preparation and Use of Nuclease-Treated Rabbit Reticulocyte Lysates for the Translation of Eukaryotic Messenger RNA. In Methods in Enzymology; Methods in Enzymology; Elsevier, 1983; pp 50–74. 10.1016/s0076-6879(83)96008-1. [DOI] [PubMed] [Google Scholar]
- Anderson C. W.; Straus J. W.; Dudock B. S.. Preparation of a Cell-Free Protein-Synthesizing System from Wheat Germ. In Recombinant DNA Methodology; Elsevier, 1989; pp 677–685. 10.1016/b978-0-12-765560-4.50044-7. [DOI] [PubMed] [Google Scholar]
- Moore S. J.; MacDonald J. T.; Wienecke S.; Ishwarbhai A.; Tsipa A.; Aw R.; Kylilis N.; Bell D. J.; McClymont D. W.; Jensen K.; Polizzi K. M.; Biedendieck R.; Freemont P. S. Rapid Acquisition and Model-Based Analysis of Cell-Free Transcription-Translation Reactions from Nonmodel Bacteria. Proc. Natl. Acad. Sci. U. S. A. 2018, 115 (19), E4340-E4349 10.1073/pnas.1715806115. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kelwick R.; Webb A. J.; MacDonald J. T.; Freemont P. S. Development of a Bacillus Subtilis Cell-Free Transcription-Translation System for Prototyping Regulatory Elements. Metab. Eng. 2016, 38, 370–381. 10.1016/j.ymben.2016.09.008. [DOI] [PubMed] [Google Scholar]
- Zhang L.; Lin X.; Wang T.; Guo W.; Lu Y. Development and Comparison of Cell-Free Protein Synthesis Systems Derived from Typical Bacterial Chassis. Bioresour. Bioprocess. 2021, 8 (1), 58. 10.1186/s40643-021-00413-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Krüger A.; Mueller A. P.; Rybnicky G. A.; Engle N. L.; Yang Z. K.; Tschaplinski T. J.; Simpson S. D.; Köpke M.; Jewett M. C. Development of a Clostridia-Based Cell-Free System for Prototyping Genetic Parts and Metabolic Pathways. Metab. Eng. 2020, 62, 95–105. 10.1016/j.ymben.2020.06.004. [DOI] [PubMed] [Google Scholar]
- Fábrega M.-J.; Knödlseder N.; Nevot G.; Sanvicente M.; Toloza L.; Santos-Moreno J.; Güell M. Establishing a Cell-Free Transcription-Translation Platform for Cutibacterium Acnes to Prototype Engineered Metabolic and Synthetic Biology. ACS Biomater. Sci. Eng. 2023, 9, 5101. 10.1021/acsbiomaterials.1c00894. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yim S. S.; Johns N. I.; Park J.; Gomes A. L.; McBee R. M.; Richardson M.; Ronda C.; Chen S. P.; Garenne D.; Noireaux V.; Wang H. H. Multiplex Transcriptional Characterizations across Diverse Bacterial Species Using Cell-Free Systems. Mol. Syst. Biol. 2019, 15 (8), e8875 10.15252/msb.20198875. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nakashima N.; Tamura T. Cell-Free Protein Synthesis Using Cell Extract of Pseudomonas Fluorescens and CspA Promoter. Biochem. Biophys. Res. Commun. 2004, 319 (2), 671–676. 10.1016/j.bbrc.2004.05.034. [DOI] [PubMed] [Google Scholar]
- Wang H.; Li J.; Jewett M. C. Development of a Pseudomonas Putida Cell-Free Protein Synthesis Platform for Rapid Screening of Gene Regulatory Elements. Synth. Biol. 2018, 3 (1), ysy003. 10.1093/synbio/ysy003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li J.; Wang H.; Kwon Y.-C.; Jewett M. C. Establishing a High Yieldingstreptomyces-Based Cell-Free Protein Synthesis System. Biotechnol. Bioeng. 2017, 114 (6), 1343–1353. 10.1002/bit.26253. [DOI] [PubMed] [Google Scholar]
- Xu H.; Yang C.; Tian X.; Chen Y.; Liu W.-Q.; Li J. Regulatory Part Engineering for High-Yield Protein Synthesis in an All-Streptomyces-Based Cell-Free Expression System. ACS Synth. Biol. 2022, 11 (2), 570–578. 10.1021/acssynbio.1c00587. [DOI] [PubMed] [Google Scholar]
- Xu H.; Liu W.-Q.; Li J. Translation Related Factors Improve the Productivity of a Streptomyces-Based Cell-Free Protein Synthesis System. ACS Synth. Biol. 2020, 9 (5), 1221–1224. 10.1021/acssynbio.0c00140. [DOI] [PubMed] [Google Scholar]
- Moore S. J.; Lai H.-E.; Chee S.-M.; Toh M.; Coode S.; Chengan K.; Capel P.; Corre C.; de Los Santos E. L.; Freemont P. S. A Streptomyces Venezuelae Cell-Free Toolkit for Synthetic Biology. ACS Synth. Biol. 2021, 10 (2), 402–411. 10.1021/acssynbio.0c00581. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Endoh T.; Kanai T.; Imanaka T. A Highly Productive System for Cell-Free Protein Synthesis Using a Lysate of the Hyperthermophilic Archaeon, Thermococcus Kodakaraensis. Appl. Microbiol. Biotechnol. 2007, 74 (5), 1153–1161. 10.1007/s00253-006-0753-3. [DOI] [PubMed] [Google Scholar]
- Des Soye B. J.; Davidson S. R.; Weinstock M. T.; Gibson D. G.; Jewett M. C. Establishing a High-Yielding Cell-Free Protein Synthesis Platform Derived from Vibrio Natriegens. ACS Synth. Biol. 2018, 7 (9), 2245–2255. 10.1021/acssynbio.8b00252. [DOI] [PubMed] [Google Scholar]
- McDonald N. D.; Rhea K. A.; Berk K. L.; Zacharko J. L.; Miklos A. E. Cell-Free Protein Systems from Yersinia Pestis Are Functional and Growth-Temperature Dependent. ACS Synth. Biol. 2021, 10 (12), 3604–3607. 10.1021/acssynbio.1c00505. [DOI] [PubMed] [Google Scholar]
- Ito M. Development and Implementation of Silkworm Cell-Free Protein Synthesis Systems. J. Biol. Macromol. 2021, 21 (1), 3–11. 10.14533/jbm.21.3. [DOI] [Google Scholar]
- Thoring L.; Dondapati S. K.; Stech M.; Wüstenhagen D. A.; Kubick S. High-Yield Production of “Difficult-to-Express” Proteins in a Continuous Exchange Cell-Free System Based on CHO Cell Lysates. Sci. Rep. 2017, 7 (1), 11710. 10.1038/s41598-017-12188-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brödel A. K.; Sonnabend A.; Roberts L. O.; Stech M.; Wüstenhagen D. A.; Kubick S. IRES-Mediated Translation of Membrane Proteins and Glycoproteins in Eukaryotic Cell-Free Systems. PLoS One 2013, 8 (12), e82234 10.1371/journal.pone.0082234. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mikami S.; Masutani M.; Sonenberg N.; Yokoyama S.; Imataka H. An Efficient Mammalian Cell-Free Translation System Supplemented with Translation Factors. Protein Expr. Purif. 2006, 46 (2), 348–357. 10.1016/j.pep.2005.09.021. [DOI] [PubMed] [Google Scholar]
- Mureev S.; Kovtun O.; Nguyen U. T. T.; Alexandrov K. Species-Independent Translational Leaders Facilitate Cell-Free Expression. Nat. Biotechnol. 2009, 27 (8), 747–752. 10.1038/nbt.1556. [DOI] [PubMed] [Google Scholar]
- Gupta M. D.; Flaskamp Y.; Roentgen R.; Juergens H.; Armero-Gimenez J.; Albrecht F.; Hemmerich J.; Arfi Z. A.; Neuser J.; Spiegel H.; Schillberg S.; Yeliseev A.; Song L.; Qiu J.; Williams C.; Finnern R. Scaling Eukaryotic Cell-free Protein Synthesis Achieved with the Versatile and High-yielding Tobacco BY-2 Cell Lysate. Biotechnol. Bioeng. 2023, 120 (10), 2890–2906. 10.1002/bit.28461. [DOI] [PubMed] [Google Scholar]
- Spice A. J.; Aw R.; Bracewell D. G.; Polizzi K. M. Improving the Reaction Mix of a Pichia Pastoris Cell-Free System Using a Design of Experiments Approach to Minimise Experimental Effort. Synth. Syst. Biotechnol. 2020, 5 (3), 137–144. 10.1016/j.synbio.2020.06.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schoborg J. A.; Clark L. G.; Choudhury A.; Hodgman C. E.; Jewett M. C. Yeast Knockout Library Allows for Efficient Testing of Genomic Mutations for Cell-Free Protein Synthesis. Synth. Syst. Biotechnol. 2016, 1 (1), 2–6. 10.1016/j.synbio.2016.02.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schoborg J. A.; Hodgman C. E.; Anderson M. J.; Jewett M. C. Substrate Replenishment and Byproduct Removal Improve Yeast Cell-Free Protein Synthesis. Biotechnol. J. 2014, 9 (5), 630–640. 10.1002/biot.201300383. [DOI] [PubMed] [Google Scholar]
- Merk H.; Rues R.-B.; Gless C.; Beyer K.; Dong F.; Dötsch V.; Gerrits M.; Bernhard F. Biosynthesis of Membrane Dependent Proteins in Insect Cell Lysates: Identification of Limiting Parameters for Folding and Processing. Biol. Chem. 2015, 396 (9–10), 1097–1107. 10.1515/hsz-2015-0105. [DOI] [PubMed] [Google Scholar]
- Ezure T.; Suzuki T.; Higashide S.; Shintani E.; Endo K.; Kobayashi S.-I.; Shikata M.; Ito M.; Tanimizu K.; Nishimura O. Cell-Free Protein Synthesis System Prepared from Insect Cells by Freeze-Thawing. Biotechnol. Prog. 2006, 22 (6), 1570–1577. 10.1021/bp060110v. [DOI] [PubMed] [Google Scholar]
- Beebe E. T.; Makino S.-I.; Nozawa A.; Matsubara Y.; Frederick R. O.; Primm J. G.; Goren M. A.; Fox B. G. Robotic Large-Scale Application of Wheat Cell-Free Translation to Structural Studies Including Membrane Proteins. N. Biotechnol. 2011, 28 (3), 239–249. 10.1016/j.nbt.2010.07.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Harbers M. Wheat Germ Systems for Cell-Free Protein Expression. FEBS Lett. 2014, 588 (17), 2762–2773. 10.1016/j.febslet.2014.05.061. [DOI] [PubMed] [Google Scholar]
- Sridharan H.; Piorino F.; Styczynski M. P. Systems Biology-Based Analysis of Cell-Free Systems. Curr. Opin. Biotechnol. 2022, 75, 102703. 10.1016/j.copbio.2022.102703. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Garenne D.; Beisel C. L.; Noireaux V. Characterization of the All-E. Coli Transcription-Translation System MyTXTL by Mass Spectrometry. Rapid Commun. Mass Spectrom. 2019, 33 (11), 1036–1048. 10.1002/rcm.8438. [DOI] [PubMed] [Google Scholar]
- Foshag D.; Henrich E.; Hiller E.; Schäfer M.; Kerger C.; Burger-Kentischer A.; Diaz-Moreno I.; García-Mauriño S. M.; Dötsch V.; Rupp S.; Bernhard F. The E. Coli S30 Lysate Proteome: A Prototype for Cell-Free Protein Production. N. Biotechnol. 2018, 40, 245–260. 10.1016/j.nbt.2017.09.005. [DOI] [PubMed] [Google Scholar]
- Hurst G. B.; Asano K. G.; Doktycz C. J.; Consoli E. J.; Doktycz W. L.; Foster C. M.; Morrell-Falvey J. L.; Standaert R. F.; Doktycz M. J. Proteomics-Based Tools for Evaluation of Cell-Free Protein Synthesis. Anal. Chem. 2017, 89 (21), 11443–11451. 10.1021/acs.analchem.7b02555. [DOI] [PubMed] [Google Scholar]
- Dinglasan J. L. N.; Reeves D. T.; Hettich R. L.; Doktycz M. J.. Liquid Chromatography Coupled to Refractive Index or Mass Spectrometric Detection for Metabolite Profiling in Lysate-Based Cell-Free Systems. J. Vis. Exp. 2021, No. (175), . 10.3791/62852. [DOI] [PubMed] [Google Scholar]
- Mohr B.; Giannone R. J.; Hettich R. L.; Doktycz M. J. Targeted Growth Medium Dropouts Promote Aromatic Compound Synthesis in Crude E. Coli Cell-Free Systems. ACS Synth. Biol. 2020, 9 (11), 2986–2997. 10.1021/acssynbio.9b00524. [DOI] [PubMed] [Google Scholar]
- Contreras-Llano L. E.; Meyer C.; Liu Y.; Sarker M.; Lim S.; Longo M. L.; Tan C. Holistic Engineering of Cell-Free Systems through Proteome-Reprogramming Synthetic Circuits. Nat. Commun. 2020, 11 (1), 3138. 10.1038/s41467-020-16900-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Falgenhauer E.; von Schönberg S.; Meng C.; Mückl A.; Vogele K.; Emslander Q.; Ludwig C.; Simmel F. C. Evaluation of an E. Coli Cell Extract Prepared by Lysozyme-Assisted Sonication via Gene Expression, Phage Assembly and Proteomics. Chembiochem 2021, 22 (18), 2805–2813. 10.1002/cbic.202100257. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Levine M. Z.; So B.; Mullin A. C.; Fanter R.; Dillard K.; Watts K. R.; La Frano M. R.; Oza J. P. Activation of Energy Metabolism through Growth Media Reformulation Enables a 24-h Workflow for Cell-Free Expression. ACS Synth. Biol. 2020, 9 (10), 2765–2774. 10.1021/acssynbio.0c00283. [DOI] [PubMed] [Google Scholar]
- Miguez A. M.; McNerney M. P.; Styczynski M. P. Metabolic Profiling of Escherichia Coli-Based Cell-Free Expression Systems for Process Optimization. Ind. Eng. Chem. Res. 2019, 58 (50), 22472–22482. 10.1021/acs.iecr.9b03565. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miguez A. M.; Zhang Y.; Piorino F.; Styczynski M. P. Metabolic Dynamics in Escherichia Coli-Based Cell-Free Systems. ACS Synth. Biol. 2021, 10 (9), 2252–2265. 10.1021/acssynbio.1c00167. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Campo A. M.; Raig R.; Hershewe J. M.. Peeking inside the Black Box: NMR Metabolomics for Optimizing Cell-Free Protein Synthesis. In Women in Aerospace Materials; Springer International Publishing: Cham, 2020; pp 19–33. 10.1007/978-3-030-40779-7_2. [DOI] [Google Scholar]
- Moore S. J.; Lai H.-E.; Li J.; Freemont P. S. Streptomyces Cell-Free Systems for Natural Product Discovery and Engineering. Nat. Prod. Rep. 2023, 40 (2), 228–236. 10.1039/D2NP00057A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Failmezger J.; Scholz S.; Blombach B.; Siemann-Herzberg M.. Cell-Free Protein Synthesis From Fast-Growing Vibrio Natriegens. Front. Microbiol. 2018, 9. 10.3389/fmicb.2018.01146. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wiegand D. J.; Lee H. H.; Ostrov N.; Church G. M. Establishing a Cell-Free Vibrio Natriegens Expression System. ACS Synth. Biol. 2018, 7 (10), 2475–2479. 10.1021/acssynbio.8b00222. [DOI] [PubMed] [Google Scholar]
- Novikova I. V.; Sharma N.; Moser T.; Sontag R.; Liu Y.; Collazo M. J.; Cascio D.; Shokuhfar T.; Hellmann H.; Knoblauch M.; Evans J. E. Protein Structural Biology Using Cell-Free Platform from Wheat Germ. Adv. Struct. Chem. Imaging 2018, 4 (1), 13. 10.1186/s40679-018-0062-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dudley Q. M.; Cai Y.-M.; Kallam K.; Debreyne H.; Carrasco Lopez J. A.; Patron N. J. Biofoundry-Assisted Expression and Characterization of Plant Proteins. Synth. Biol. 2021, 6 (1), ysab029. 10.1093/synbio/ysab029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kanoi B. N.; Nagaoka H.; Morita M.; Tsuboi T.; Takashima E. Leveraging the Wheat Germ Cell-Free Protein Synthesis System to Accelerate Malaria Vaccine Development. Parasitol. Int. 2021, 80, 102224. 10.1016/j.parint.2020.102224. [DOI] [PubMed] [Google Scholar]
- Roberts B. E.; Paterson B. M. Efficient Translation of Tobacco Mosaic Virus RNA and Rabbit Globin 9S RNA in a Cell-Free System from Commercial Wheat Germ. Proc. Natl. Acad. Sci. U. S. A. 1973, 70 (8), 2330–2334. 10.1073/pnas.70.8.2330. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Buntru M.; Vogel S.; Spiegel H.; Schillberg S. Tobacco BY-2 Cell-Free Lysate: An Alternative and Highly-Productive Plant-Based in Vitro Translation System. BMC Biotechnol. 2014, 14 (1), 37. 10.1186/1472-6750-14-37. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Buntru M.; Vogel S.; Stoff K.; Spiegel H.; Schillberg S. A Versatile Coupled Cell-Free Transcription-Translation System Based on Tobacco BY-2 Cell Lysates. Biotechnol. Bioeng. 2015, 112 (5), 867–878. 10.1002/bit.25502. [DOI] [PubMed] [Google Scholar]
- Armero-Gimenez J.; Wilbers R.; Schots A.; Williams C.; Finnern R.. Rapid Screening and Scaled Manufacture of Immunogenic Virus-like Particles in a Tobacco BY-2 Cell-Free Protein Synthesis System. Front. Immunol. 2023, 14. 10.3389/fimmu.2023.1088852. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Spencer D.; Wildman S. G. The Incorporation of Amino Acids into Protein by Cell-Free Extracts from Tobacco Leaves. Biochemistry 1964, 3 (7), 954–959. 10.1021/bi00895a019. [DOI] [PubMed] [Google Scholar]
- Yukawa M.; Kuroda H.; Sugiura M. A New in Vitro Translation System for Non-Radioactive Assay from Tobacco Chloroplasts: Effect of Pre-MRNA Processing on Translation in Vitro. Plant J. 2007, 49 (2), 367–376. 10.1111/j.1365-313X.2006.02948.x. [DOI] [PubMed] [Google Scholar]
- Murota K.; Hagiwara-Komoda Y.; Komoda K.; Onouchi H.; Ishikawa M.; Naito S. Arabidopsis Cell-Free Extract, ACE, a New in Vitro Translation System Derived from Arabidopsis Callus Cultures. Plant Cell Physiol. 2011, 52 (8), 1443–1453. 10.1093/pcp/pcr080. [DOI] [PubMed] [Google Scholar]
- Kasai K.; Kanno T.; Endo Y.; Wakasa K.; Tozawa Y. Guanosine Tetra- and Pentaphosphate Synthase Activity in Chloroplasts of a Higher Plant: Association with 70S Ribosomes and Inhibition by Tetracycline. Nucleic Acids Res. 2004, 32 (19), 5732–5741. 10.1093/nar/gkh916. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rilkey G. J. P. Effects of High Temperature on Protein Synthesis during Germination of Maize (Zea Mays L.). Planta 1981, 151 (1), 75–80. 10.1007/BF00384240. [DOI] [PubMed] [Google Scholar]
- Faust F.; Schubert S. In Vitro Protein Synthesis of Sugar Beet (Beta Vulgaris) and Maize (Zea Mays) Is Differentially Inhibited When Potassium Is Substituted by Sodium. Plant Physiol. Biochem. 2017, 118, 228–234. 10.1016/j.plaphy.2017.06.016. [DOI] [PubMed] [Google Scholar]
- Martin R. W.; Majewska N. I.; Chen C. X.; Albanetti T. E.; Jimenez R. B. C.; Schmelzer A. E.; Jewett M. C.; Roy V. Development of a CHO-Based Cell-Free Platform for Synthesis of Active Monoclonal Antibodies. ACS Synth. Biol. 2017, 6 (7), 1370–1379. 10.1021/acssynbio.7b00001. [DOI] [PubMed] [Google Scholar]
- Stech M.; Nikolaeva O.; Thoring L.; Stöcklein W. F. M.; Wüstenhagen D. A.; Hust M.; Dübel S.; Kubick S. Cell-Free Synthesis of Functional Antibodies Using a Coupled in Vitro Transcription-Translation System Based on CHO Cell Lysates. Sci. Rep. 2017, 7 (1), 12030. 10.1038/s41598-017-12364-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mikami S.; Kobayashi T.; Yokoyama S.; Imataka H. A Hybridoma-Based in Vitro Translation System That Efficiently Synthesizes Glycoproteins. J. Biotechnol. 2006, 127 (1), 65–78. 10.1016/j.jbiotec.2006.06.018. [DOI] [PubMed] [Google Scholar]
- Shibutani M.; Kim E.; Lazarovici P.; Oshima M.; Guroff G. Preparation of a Cell-Free Translation System from PC12 Cells. Neurochem. Res. 1996, 21 (7), 801–807. 10.1007/BF02532303. [DOI] [PubMed] [Google Scholar]
- Morley S. J.; Buhl W.-J.; Jackson R. J. A Rabbit Reticulocyte Factor Which Stimulates Protein Synthesis in Several Mammalian Cell-Free Systems. Biochim. Biophys. Acta Gene Struct. Expression 1985, 825 (1), 57–69. 10.1016/0167-4781(85)90079-X. [DOI] [PubMed] [Google Scholar]
- Kopniczky M. B.; Canavan C.; McClymont D. W.; Crone M. A.; Suckling L.; Goetzmann B.; Siciliano V.; MacDonald J. T.; Jensen K.; Freemont P. S. Cell-Free Protein Synthesis as a Prototyping Platform for Mammalian Synthetic Biology. ACS Synth. Biol. 2020, 9 (1), 144–156. 10.1021/acssynbio.9b00437. [DOI] [PubMed] [Google Scholar]
- Wang Y.; Wang T.; Chen X.; Lu Y.. IRES-Mediated Pichia Pastoris Cell-Free Protein Synthesis. Bioresour. Bioprocess. 2023, 10 ( (1), ). 10.1186/s40643-023-00653-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aw R.; Spice A. J.; Polizzi K. M.. Methods for Expression of Recombinant Proteins Using a Pichia Pastoris Cell-free System. Curr. Protoc. Protein Sci. 2020, 102 ( (1), ). 10.1002/cpps.115. [DOI] [PubMed] [Google Scholar]
- Gagoski D.; Polinkovsky M. E.; Mureev S.; Kunert A.; Johnston W.; Gambin Y.; Alexandrov K. Performance Benchmarking of Four Cell-free Protein Expression Systems. Biotechnol. Bioeng. 2016, 113 (2), 292–300. 10.1002/bit.25814. [DOI] [PubMed] [Google Scholar]
- Hunt A. C.; Vögeli B.; Hassan A. O.; Guerrero L.; Kightlinger W.; Yoesep D. J.; Krüger A.; DeWinter M.; Diamond M. S.; Karim A. S.; Jewett M. C.. A Rapid Cell-Free Expression and Screening Platform for Antibody Discovery. Nat. Commun. 2023, 14 ( (1), ). 10.1038/s41467-023-38965-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martin R. W.; Des Soye B. J.; Kwon Y. C.; Kay J.; Davis R. G.; Thomas P. M.; Majewska N. I.; Chen C. X.; Marcum R. D.; Weiss M. G.; Stoddart A. E.; Amiram M.; Ranji Charna A. K.; Patel J. R.; Isaacs F. J.; Kelleher N. L.; Hong S. H.; Jewett M. C. Cell-Free Protein Synthesis from Genomically Recoded Bacteria Enables Multisite Incorporation of Noncanonical Amino Acids. Nat. Commun. 2018, 9 (1), 1–9. 10.1038/s41467-018-03469-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sivashanmugam A.; Murray V.; Cui C.; Zhang Y.; Wang J.; Li Q. Practical Protocols for Production of Very High Yields of Recombinant Proteins Using Escherichia Coli. Protein Sci. 2009, 18 (5), 936–948. 10.1002/pro.102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Taylor T.; Denson J.-P.; Esposito D.. Optimizing Expression and Solubility of Proteins in E. Coli Using Modified Media and Induction Parameters. In Methods Mol. Biol.; Methods in molecular biology (Clifton, N.J.); Springer New York: New York, NY, 2017; pp 65–82. 10.1007/978-1-4939-6887-9_5. [DOI] [PubMed] [Google Scholar]
- Jewett M. C.; Swartz J. R. Substrate Replenishment Extends Protein Synthesis with an in Vitro Translation System Designed to Mimic the Cytoplasm. Biotechnol. Bioeng. 2004, 87 (4), 465–472. 10.1002/bit.20139. [DOI] [PubMed] [Google Scholar]
- Kim R. G.; Choi C. Y. Expression-Independent Consumption of Substrates in Cell-Free Expression System from Escherichia Coli. J. Biotechnol. 2000, 84 (1), 27–32. 10.1016/S0168-1656(00)00326-6. [DOI] [PubMed] [Google Scholar]
- Silverman A. D.; Kelley-Loughnane N.; Lucks J. B.; Jewett M. C. Deconstructing Cell-Free Extract Preparation for in Vitro Activation of Transcriptional Genetic Circuitry. ACS Synth. Biol. 2019, 8 (2), 403–414. 10.1021/acssynbio.8b00430. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mullin A. C.; Slouka T.; Oza J. P. Simple Extract Preparation Methods for E. Coli-Based Cell-Free Expression. Methods Mol. Biol. 2022, 2433, 51–64. 10.1007/978-1-0716-1998-8_2. [DOI] [PubMed] [Google Scholar]
- Hunt J. P.; Zhao E. L.; Soltani M.; Frei M.; Nelson J. A. D.; Bundy B. C. Streamlining the Preparation of “Endotoxin-Free” ClearColi Cell Extract with Autoinduction Media for Cell-Free Protein Synthesis of the Therapeutic Protein Crisantaspase. Synth. Syst. Biotechnol. 2019, 4 (4), 220–224. 10.1016/j.synbio.2019.11.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kwon Y.-C.; Jewett M. C. High-Throughput Preparation Methods of Crude Extract for Robust Cell-Free Protein Synthesis. Sci. Rep. 2015, 5 (1), 8663. 10.1038/srep08663. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Failmezger J.; Rauter M.; Nitschel R.; Kraml M.; Siemann-Herzberg M. Cell-Free Protein Synthesis from Non-Growing, Stressed Escherichia Coli. Sci. Rep. 2017, 7 (1), 16524. 10.1038/s41598-017-16767-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dopp J. L.; Jo Y. R.; Reuel N. F. Methods to Reduce Variability in E. Coli-Based Cell-Free Protein Expression Experiments. Synth. Syst. Biotechnol. 2019, 4 (4), 204–211. 10.1016/j.synbio.2019.10.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang W. C.; Patel K. G.; Wong H. E.; Swartz J. R. Simplifying and Streamlining Escherichia Coli-Based Cell-Free Protein Synthesis. Biotechnol. Prog. 2012, 28 (2), 413–420. 10.1002/btpr.1509. [DOI] [PubMed] [Google Scholar]
- Liu D. V.; Zawada J. F.; Swartz J. R. Streamlining Escherichia Coli S30 Extract Preparation for Economical Cell-Free Protein Synthesis. Biotechnol. Prog. 2005, 21 (2), 460–465. 10.1021/bp049789y. [DOI] [PubMed] [Google Scholar]
- Kim T.-W.; Keum J.-W.; Oh I.-S.; Choi C.-Y.; Park C.-G.; Kim D.-M. Simple Procedures for the Construction of a Robust and Cost-Effective Cell-Free Protein Synthesis System. J. Biotechnol. 2006, 126 (4), 554–561. 10.1016/j.jbiotec.2006.05.014. [DOI] [PubMed] [Google Scholar]
- Shrestha P.; Holland T. M.; Bundy B. C. Streamlined Extract Preparation for Escherichia Coli-Based Cell-Free Protein Synthesis by Sonication or Bead Vortex Mixing. Biotechniques 2012, 53 (3), 163–174. 10.2144/0000113924. [DOI] [PubMed] [Google Scholar]
- Sun Z. Z.; Hayes C. A.; Shin J.; Caschera F.; Murray R. M.; Noireaux V. Protocols for Implementing an Escherichia Coli Based TX-TL Cell-Free Expression System for Synthetic Biology. J. Vis. Exp. 2013, (79), e50762 10.3791/50762. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fujiwara K.; Doi N. Biochemical Preparation of Cell Extract for Cell-Free Protein Synthesis without Physical Disruption. PLoS One 2016, 11 (4), e0154614 10.1371/journal.pone.0154614. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Didovyk A.; Tonooka T.; Tsimring L.; Hasty J. Rapid and Scalable Preparation of Bacterial Lysates for Cell-Free Gene Expression. ACS Synth. Biol. 2017, 6 (12), 2198–2208. 10.1021/acssynbio.7b00253. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hershewe J. M.; Warfel K. F.; Iyer S. M.; Peruzzi J. A.; Sullivan C. J.; Roth E. W.; DeLisa M. P.; Kamat N. P.; Jewett M. C. Improving Cell-Free Glycoprotein Synthesis by Characterizing and Enriching Native Membrane Vesicles. Nat. Commun. 2021, 12 (1), 2363. 10.1038/s41467-021-22329-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nirenberg M. W.[3] Cell-Free Protein Synthesis Directed by Messenger RNA. In Methods in Enzymology; Methods in Enzymology; Elsevier, 1963; pp 17–23. 10.1016/0076-6879(63)06142-5. [DOI] [Google Scholar]
- Rasor B. J.; Chirania P.; Rybnicky G. A.; Giannone R. J.; Engle N. L.; Tschaplinski T. J.; Karim A. S.; Hettich R. L.; Jewett M. C. Mechanistic Insights into Cell-Free Gene Expression through an Integrated -Omics Analysis of Extract Processing Methods. ACS Synth. Biol. 2023, 12 (2), 405–418. 10.1021/acssynbio.2c00339. [DOI] [PubMed] [Google Scholar]
- Karig D. K.; Bessling S.; Thielen P.; Zhang S.; Wolfe J. Preservation of Protein Expression Systems at Elevated Temperatures for Portable Therapeutic Production. J. R. Soc. Interface 2017, 14 (129), 20161039. 10.1098/rsif.2016.1039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pardee K.; Slomovic S.; Nguyen P. Q.; Lee J. W.; Donghia N.; Burrill D.; Ferrante T.; McSorley F. R.; Furuta Y.; Vernet A.; Lewandowski M.; Boddy C. N.; Joshi N. S.; Collins J. J. Portable, on-Demand Biomolecular Manufacturing. Cell 2016, 167 (1), 248–259. 10.1016/j.cell.2016.09.013. [DOI] [PubMed] [Google Scholar]
- Salehi A. S. M.; Smith M. T.; Bennett A. M.; Williams J. B.; Pitt W. G.; Bundy B. C. Cell-Free Protein Synthesis of a Cytotoxic Cancer Therapeutic: Onconase Production and a Just-Add-Water Cell-Free System. Biotechnol. J. 2016, 11 (2), 274–281. 10.1002/biot.201500237. [DOI] [PubMed] [Google Scholar]
- Stark J. C.; Jaroentomeechai T.; Moeller T. D.; Hershewe J. M.; Warfel K. F.; Moricz B. S.; Martini A. M.; Dubner R. S.; Hsu K. J.; Stevenson T. C.; Jones B. D.; DeLisa M. P.; Jewett M. C. On-Demand Biomanufacturing of Protective Conjugate Vaccines. Science Advances 2021, 7 (6), eabe9444 10.1126/sciadv.abe9444. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu X.; Silverman A. D.; Alam K. K.; Iverson E.; Lucks J. B.; Jewett M. C.; Raman S. Design of a Transcriptional Biosensor for the Portable, on-Demand Detection of Cyanuric Acid. ACS Synth. Biol. 2020, 9 (1), 84–94. 10.1021/acssynbio.9b00348. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thavarajah W.; Silverman A. D.; Verosloff M. S.; Kelley-Loughnane N.; Jewett M. C.; Lucks J. B. Point-of-Use Detection of Environmental Fluoride via a Cell-Free Riboswitch-Based Biosensor. ACS Synth. Biol. 2020, 9 (1), 10–18. 10.1021/acssynbio.9b00347. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang A.; Nguyen P. Q.; Stark J. C.; Takahashi M. K.; Donghia N.; Ferrante T.; Dy A. J.; Hsu K. J.; Dubner R. S.; Pardee K.; Jewett M. C.; Collins J. J. BioBitsTM Explorer: A Modular Synthetic Biology Education Kit. Sci. Adv. 2018, 4 (8), eaat5105 10.1126/sciadv.aat5105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stark J. C.; Huang A.; Nguyen P. Q.; Dubner R. S.; Hsu K. J.; Ferrante T. C.; Anderson M.; Kanapskyte A.; Mucha Q.; Packett J. S.; Patel P.; Patel R.; Qaq D.; Zondor T.; Burke J.; Martinez T.; Miller-Berry A.; Puppala A.; Reichert K.; Schmid M.; Brand L.; Hill L. R.; Chellaswamy J. F.; Faheem N.; Fetherling S.; Gong E.; Gonzalzles E. M.; Granito T.; Koritsaris J.; Nguyen B.; Ottman S.; Palffy C.; Patel A.; Skweres S.; Slaton A.; Woods T. R.; Donghia N.; Pardee K.; Collins J. J.; Jewett M. C. BioBitsTM Bright: A Fluorescent Synthetic Biology Education Kit. Science Advances 2018, 4 (8), 1–11. 10.1126/sciadv.aat5107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smith M. T.; Berkheimer S. D.; Werner C. J.; Bundy B. C. Lyophilized Escherichia Coli-Based Cell-Free Systems for Robust, High-Density, Long-Term Storage. Biotechniques 2014, 56 (4), 186–193. 10.2144/000114158. [DOI] [PubMed] [Google Scholar]
- Gregorio N. E.; Kao W. Y.; Williams L. C.; Hight C. M.; Patel P.; Watts K. R.; Oza J. P. Unlocking Applications of Cell-Free Biotechnology through Enhanced Shelf Life and Productivity of E. Coli Extracts. ACS Synth. Biol. 2020, 9 (4), 766–778. 10.1021/acssynbio.9b00433. [DOI] [PubMed] [Google Scholar]
- Guzman-Chavez F.; Arce A.; Adhikari A.; Vadhin S.; Pedroza-Garcia J. A.; Gandini C.; Ajioka J. W.; Molloy J.; Sanchez-Nieto S.; Varner J. D.; Federici F.; Haseloff J. Constructing Cell-Free Expression Systems for Low-Cost Access. ACS Synth. Biol. 2022, 11 (3), 1114–1128. 10.1021/acssynbio.1c00342. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hodgman C. E.; Jewett M. C. Optimized Extract Preparation Methods and Reaction Conditions for Improved Yeast Cell-Free Protein Synthesis. Biotechnol. Bioeng. 2013, 110 (10), 2643–2654. 10.1002/bit.24942. [DOI] [PubMed] [Google Scholar]
- PRATT J. M. Coupled Transcription-Translation in Prokaryotic Cell-Free Systems. Transcription and translation: a practical approach. 1984, 179–209. [Google Scholar]
- Kim D. M.; Swartz J. R. Regeneration of Adenosine Triphosphate from Glycolytic Intermediates for Cell-Free Protein Synthesis. Biotechnol. Bioeng. 2001, 74 (4), 309–316. 10.1002/bit.1121. [DOI] [PubMed] [Google Scholar]
- Anderson C. W.; Straus J. W.; Dudock B. S. Preparation of a Cell-Free Protein-Synthesizing System from Wheat Germ. Methods Enzymol. 1983, 101, 635–644. 10.1016/0076-6879(83)01044-7. [DOI] [PubMed] [Google Scholar]
- Spirin A. S.; Baranov V. I.; Ryabova L. A.; Ovodov S. Y.; Alakhov Y. B. A Continuous Cell-Free Translation System Capable of Producing Polypeptides in High Yield. Science 1988, 242 (4882), 1162–1164. 10.1126/science.3055301. [DOI] [PubMed] [Google Scholar]
- Ryabova L. A.; Vinokurov L. M.; Shekhovtsova E. A.; Alakhov Y. B.; Spirin A. S. Acetyl Phosphate as an Energy Source for Bacterial Cell-Free Translation Systems. Anal. Biochem. 1995, 226 (1), 184–186. 10.1006/abio.1995.1208. [DOI] [PubMed] [Google Scholar]
- Kim T.-W.; Keum J.-W.; Oh I.-S.; Choi C.-Y.; Kim H.-C.; Kim D.-M. An Economical and Highly Productive Cell-Free Protein Synthesis System Utilizing Fructose-1,6-Bisphosphate as an Energy Source. J. Biotechnol. 2007, 130, 389–393. 10.1016/j.jbiotec.2007.05.002. [DOI] [PubMed] [Google Scholar]
- Calhoun K. A.; Swartz J. R. An Economical Method for Cell-Free Protein Synthesis Using Glucose and Nucleoside Monophosphates. Biotechnol. Prog. 2005, 21 (4), 1146–1153. 10.1021/bp050052y. [DOI] [PubMed] [Google Scholar]
- Cai Q.; Hanson J. A.; Steiner A. R.; Tran C.; Masikat M. R.; Chen R.; Zawada J. F.; Sato A. K.; Hallam T. J.; Yin G. A Simplified and Robust Protocol for Immunoglobulin Expression in Escherichia Coli Cell-Free Protein Synthesis Systems. Biotechnol. Prog. 2015, 31 (3), 823–831. 10.1002/btpr.2082. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kim D. M.; Swartz J. R. Prolonging Cell-Free Protein Synthesis by Selective Reagent Additions. Biotechnol. Prog. 2000, 16 (3), 385–390. 10.1021/bp000031y. [DOI] [PubMed] [Google Scholar]
- Kim D. M.; Swartz J. R. Prolonging Cell-Free Protein Synthesis with a Novel ATP Regeneration System. Biotechnol. Bioeng. 1999, 66 (3), 180–188. . [DOI] [PubMed] [Google Scholar]
- Sitaraman K.; Esposito D.; Klarmann G.; Le Grice S. F.; Hartley J. L.; Chatterjee D. K. A Novel Cell-Free Protein Synthesis System. J. Biotechnol. 2004, 110 (3), 257–263. 10.1016/j.jbiotec.2004.02.014. [DOI] [PubMed] [Google Scholar]
- Caschera F.; Noireaux V. A Cost-Effective Polyphosphate-Based Metabolism Fuels an All E. Coli Cell-Free Expression System. Metab. Eng. 2015, 27, 29–37. 10.1016/j.ymben.2014.10.007. [DOI] [PubMed] [Google Scholar]
- Wang Y.; Zhang Y.-H. P. Cell-Free Protein Synthesis Energized by Slowly-Metabolized Maltodextrin. BMC Biotechnol. 2009, 9, 58. 10.1186/1472-6750-9-58. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Caschera F.; Noireaux V. Synthesis of 2.3 Mg/Ml of Protein with an All Escherichia Coli Cell-Free Transcription-Translation System. Biochimie 2014, 99, 162–168. 10.1016/j.biochi.2013.11.025. [DOI] [PubMed] [Google Scholar]
- Kim D.-M.; Swartz J. R. Oxalate Improves Protein Synthesis by Enhancing ATP Supply in a Cell-Free System Derived from Escherichia Coli. Biotechnol. Lett. 2000, 22 (19), 1537–1542. 10.1023/A:1005624811710. [DOI] [Google Scholar]
- Shin J.; Noireaux V. Efficient Cell-Free Expression with the Endogenous E. Coli RNA Polymerase and Sigma Factor 70. J. Biol. Eng. 2010, 4, 8. 10.1186/1754-1611-4-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fujiwara K.; Nomura S.-I. M. Condensation of an Additive-Free Cell Extract to Mimic the Conditions of Live Cells. PLoS One 2013, 8 (1), e54155 10.1371/journal.pone.0054155. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kai L.; Dötsch V.; Kaldenhoff R.; Bernhard F. Artificial Environments for the Co-Translational Stabilization of Cell-Free Expressed Proteins. PLoS One 2013, 8 (2), e56637 10.1371/journal.pone.0056637. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ge X.; Luo D.; Xu J. Cell-Free Protein Expression under Macromolecular Crowding Conditions. PLoS One 2011, 6 (12), e28707 10.1371/journal.pone.0028707. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pedersen A.; Hellberg K.; Enberg J.; Karlsson B. G. Rational Improvement of Cell-Free Protein Synthesis. N. Biotechnol. 2011, 28 (3), 218–224. 10.1016/j.nbt.2010.06.015. [DOI] [PubMed] [Google Scholar]
- Igarashi K.; Kashiwagi K. Polyamines: Mysterious Modulators of Cellular Functions. Biochem. Biophys. Res. Commun. 2000, 271 (3), 559–564. 10.1006/bbrc.2000.2601. [DOI] [PubMed] [Google Scholar]
- Asahara H.; Magnelli P.; Shi X.; Tuckey C.; Zhou Y.; Samuelson J. C.. Guidelines for Nucleic Acid Template Design for Optimal Cell-Free Protein Synthesis Using an Escherichia Coli Reconstituted System or a Lysate-Based System. In Recombinant Protein Expression: Prokaryotic Hosts and Cell-Free Systems; Methods in Enzymology; Elsevier, 2021; pp 351–369. 10.1016/bs.mie.2021.07.005. [DOI] [PubMed] [Google Scholar]
- Garenne D.; Bowden S.; Noireaux V. Cell-Free Expression and Synthesis of Viruses and Bacteriophages: Applications to Medicine and Nanotechnology. Curr. Opin. Syst. Biol. 2021, 28, 100373. 10.1016/j.coisb.2021.100373. [DOI] [Google Scholar]
- Jew K.; Smith P. E. J.; So B.; Kasman J.; Oza J. P.; Black M. W.. Characterizing and Improving PET Vectors for Cell-Free Expression. Front. Bioeng. Biotechnol. 2022, 10. 10.3389/fbioe.2022.895069. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kovtun O.; Mureev S.; Jung W.; Kubala M. H.; Johnston W.; Alexandrov K. Leishmania Cell-Free Protein Expression System. Methods 2011, 55 (1), 58–64. 10.1016/j.ymeth.2011.06.006. [DOI] [PubMed] [Google Scholar]
- Stech M.; Quast R. B.; Sachse R.; Schulze C.; Wüstenhagen D. A.; Kubick S. A Continuous-Exchange Cell-Free Protein Synthesis System Based on Extracts from Cultured Insect Cells. PLoS One 2014, 9 (5), e96635 10.1371/journal.pone.0096635. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Brödel A. K.; Sonnabend A.; Kubick S. Cell-free Protein Expression Based on Extracts from CHO Cells. Biotechnol. Bioeng. 2014, 111 (1), 25–36. 10.1002/bit.25013. [DOI] [PubMed] [Google Scholar]
- Zemella A.; Thoring L.; Hoffmeister C.; Kubick S. Cell-Free Protein Synthesis: Pros and Cons of Prokaryotic and Eukaryotic Systems. Chembiochem 2015, 16 (17), 2420–2431. 10.1002/cbic.201500340. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hino M.; Kataoka M.; Kajimoto K.; Yamamoto T.; Kido J.-I.; Shinohara Y.; Baba Y. Efficiency of Cell-Free Protein Synthesis Based on a Crude Cell Extract from Escherichia Coli, Wheat Germ, and Rabbit Reticulocytes. J. Biotechnol. 2008, 133 (2), 183–189. 10.1016/j.jbiotec.2007.08.008. [DOI] [PubMed] [Google Scholar]
- Underwood K. A.; Swartz J. R.; Puglisi J. D. Quantitative Polysome Analysis Identifies Limitations in Bacterial Cell-Free Protein Synthesis. Biotechnol. Bioeng. 2005, 91 (4), 425–435. 10.1002/bit.20529. [DOI] [PubMed] [Google Scholar]
- Hansen M. M. K.; Ventosa Rosquelles M.; Yelleswarapu M.; Maas R. J. M.; van Vugt-Jonker A. J.; Heus H. A.; Huck W. T. S. Protein Synthesis in Coupled and Uncoupled Cell-Free Prokaryotic Gene Expression Systems. ACS Synth. Biol. 2016, 5 (12), 1433–1440. 10.1021/acssynbio.6b00010. [DOI] [PubMed] [Google Scholar]
- McSweeney M. A.; Styczynski M. P. Effective Use of Linear DNA in Cell-Free Expression Systems. Front. Bioeng. Biotechnol. 2021, 9, 715328. 10.3389/fbioe.2021.715328. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schinn S. M.; Broadbent A.; Bradley W. T.; Bundy B. C. Protein Synthesis Directly from PCR: Progress and Applications of Cell-Free Protein Synthesis with Linear DNA. N. Biotechnol. 2016, 33 (4), 480–487. 10.1016/j.nbt.2016.04.002. [DOI] [PubMed] [Google Scholar]
- Lee K.-H.; Kim D.-M. Recent Advances in Development of Cell-Free Protein Synthesis Systems for Fast and Efficient Production of Recombinant Proteins. FEMS Microbiol. Lett. 2018, 10.1093/femsle/fny174. [DOI] [PubMed] [Google Scholar]
- Dopp J. L.; Rothstein S. M.; Mansell T. J.; Reuel N. F. Rapid Prototyping of Proteins: Mail Order Gene Fragments to Assayable Proteins within 24 h. Biotechnol. Bioeng. 2019, 116 (3), 667–676. 10.1002/bit.26912. [DOI] [PubMed] [Google Scholar]
- Sun Z. Z.; Yeung E.; Hayes C. A.; Noireaux V.; Murray R. M. Linear DNA for Rapid Prototyping of Synthetic Biological Circuits in an Escherichia Coli Based TX-TL Cell-Free System. ACS Synth. Biol. 2014, 3 (6), 387–397. 10.1021/sb400131a. [DOI] [PubMed] [Google Scholar]
- Hunt A. C.; Vögeli B.; Kightlinger W. K.; Yoesep D. J.; Krüger A.; Jewett M. C. A High-Throughput, Automated, Cell-Free Expression and Screening Platform for Antibody Discovery. bioRxiv 2021, 10.1101/2021.11.04.467378. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gagoski D.; Mureev S.; Giles N.; Johnston W.; Dahmer-Heath M.; Škalamera D.; Gonda T. J.; Alexandrov K. Gateway-Compatible Vectors for High-Throughput Protein Expression in pro- and Eukaryotic Cell-Free Systems. J. Biotechnol. 2015, 195, 1–7. 10.1016/j.jbiotec.2014.12.006. [DOI] [PubMed] [Google Scholar]
- Ojima-Kato T.; Nagai S.; Nakano H. Ecobody Technology: Rapid Monoclonal Antibody Screening Method from Single B Cells Using Cell-Free Protein Synthesis for Antigen-Binding Fragment Formation. Sci. Rep. 2017, 7 (1), 13979. 10.1038/s41598-017-14277-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hunt A. C.; Case J. B.; Park Y.-J.; Cao L.; Wu K.; Walls A. C.; Liu Z.; Bowen J. E.; Yeh H.-W.; Saini S.; Helms L.; Zhao Y. T.; Hsiang T.-Y.; Starr T. N.; Goreshnik I.; Kozodoy L.; Carter L.; Ravichandran R.; Green L. B.; Matochko W. L.; Thomson C. A.; Vögeli B.; Krüger-Gericke A.; VanBlargan L. A.; Chen R. E.; Ying B.; Bailey A. L.; Kafai N. M.; Boyken S.; Ljubetič A.; Edman N.; Ueda G.; Chow C.; Addetia A.; Panpradist N.; Gale M.; Freedman B. S.; Lutz B. R.; Bloom J. D.; Ruohola-Baker H.; Whelan S. P. J.; Stewart L.; Diamond M. S.; Veesler D.; Jewett M. C.; Baker D. Multivalent Designed Proteins Protect against SARS-CoV-2 Variants of Concern. bioRxiv 2021, 10.1101/2021.07.07.451375. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ahn J.-H.; Lee K.-H.; Shim J.-W.; Lee E. Y.; Kim D.-M. Streamlined Cell-Free Protein Synthesis from Sequence Information. Biotechnol. Bioprocess Eng. 2013, 18 (6), 1101–1108. 10.1007/s12257-013-0303-7. [DOI] [Google Scholar]
- Chiocchini C.; Vattem K.; Liss M.; Ludewig L.; Reusch T.; Rastogi I.; Webb B.; Trefzer A. From Electronic Sequence to Purified Protein Using Automated Gene Synthesis and In Vitro Transcription/Translation. ACS Synth. Biol. 2020, 9, 1714. 10.1021/acssynbio.0c00060. [DOI] [PubMed] [Google Scholar]
- Woodrow K. A.; Airen I. O.; Swartz J. R. Rapid Expression of Functional Genomic Libraries. J. Proteome Res. 2006, 5 (12), 3288–3300. 10.1021/pr050459y. [DOI] [PubMed] [Google Scholar]
- Michel-Reydellet N.; Woodrow K.; Swartz J. Increasing PCR Fragment Stability and Protein Yields in a Cell-Free System with Genetically Modified Escherichia Coli Extracts. J. Mol. Microbiol. Biotechnol. 2005, 9 (1), 26–34. 10.1159/000088143. [DOI] [PubMed] [Google Scholar]
- Ahn J.-H.; Chu H.-S.; Kim T.-W.; Oh I.-S.; Choi C.-Y.; Hahn G.-H.; Park C.-G.; Kim D.-M. Cell-Free Synthesis of Recombinant Proteins from PCR-Amplified Genes at a Comparable Productivity to That of Plasmid-Based Reactions. Biochem. Biophys. Res. Commun. 2005, 338 (3), 1346–1352. 10.1016/j.bbrc.2005.10.094. [DOI] [PubMed] [Google Scholar]
- Lesley S. A.; Brow M. A. D.; Burgess R. R. Use of in Vitro Protein Synthesis from Polymerase Chain Reaction-Generated Templates to Study Interaction of Escherichia Coli Transcription Factors with Core RNA Polymerase and for Epitope Mapping of Monoclonal Antibodies. J. Biol. Chem. 1991, 266 (4), 2632–2638. 10.1016/S0021-9258(18)52291-2. [DOI] [PubMed] [Google Scholar]
- Hadi T.; Nozzi N.; Melby J. O.; Gao W.; Fuerst D. E.; Kvam E. Rolling Circle Amplification of Synthetic DNA Accelerates Biocatalytic Determination of Enzyme Activity Relative to Conventional Methods. Sci. Rep. 2020, 10 (1), 10279. 10.1038/s41598-020-67307-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kumar G.; Chernaya G. Cell-Free Protein Synthesis Using Multiply-Primed Rolling Circle Amplification Products. Biotechniques 2009, 47 (1), 637–639. 10.2144/000113171. [DOI] [PubMed] [Google Scholar]
- Kuzminov A.; Stahl F. W. Stability of Linear DNA in RecA Mutant Escherichia Coli Cells Reflects Ongoing Chromosomal DNA Degradation. J. Bacteriol. 1997, 179 (3), 880–888. 10.1128/jb.179.3.880-888.1997. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Seki E.; Matsuda N.; Kigawa T. Multiple Inhibitory Factor Removal from an Escherichia Coli Cell Extract Improves Cell-Free Protein Synthesis. J. Biosci. Bioeng. 2009, 108 (1), 30–35. 10.1016/j.jbiosc.2009.02.011. [DOI] [PubMed] [Google Scholar]
- Batista A. C.; Levrier A.; Soudier P.; Voyvodic P. L.; Achmedov T.; Reif-Trauttmansdorff T.; DeVisch A.; Cohen-Gonsaud M.; Faulon J.-L.; Beisel C. L.; Bonnet J.; Kushwaha M. Differentially Optimized Cell-Free Buffer Enables Robust Expression from Unprotected Linear DNA in Exonuclease-Deficient Extracts. ACS Synth. Biol. 2022, 11, 732. 10.1021/acssynbio.1c00448. [DOI] [PubMed] [Google Scholar]
- Marshall R.; Maxwell C. S.; Collins S. P.; Beisel C. L.; Noireaux V. Short DNA Containing χ Sites Enhances DNA Stability and Gene Expression in E. Coli Cell-Free Transcription-Translation Systems. Biotechnol. Bioeng. 2017, 114 (9), 2137–2141. 10.1002/bit.26333. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen X.; Lu Y.. In Silico Design of Linear DNA for Robust Cell-Free Gene Expression. Front. Bioeng. Biotechnol. 2021, 9. 10.3389/fbioe.2021.670341. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shrestha P.; Smith M. T.; Bundy B. C. Cell-Free Unnatural Amino Acid Incorporation with Alternative Energy Systems and Linear Expression Templates. N. Biotechnol. 2014, 31 (1), 28–34. 10.1016/j.nbt.2013.09.002. [DOI] [PubMed] [Google Scholar]
- Zhu B.; Gan R.; Cabezas M. D.; Kojima T.; Nicol R.; Jewett M. C.; Nakano H. Increasing Cell-Free Gene Expression Yields from Linear Templates in Escherichia Coli and Vibrio Natriegens Extracts by Using DNA-Binding Proteins. Biotechnol. Bioeng. 2020, 117 (12), 3849–3857. 10.1002/bit.27538. [DOI] [PubMed] [Google Scholar]
- Norouzi M.; Panfilov S.; Pardee K. High-Efficiency Protection of Linear DNA in Cell-Free Extracts from Escherichia Coli and Vibrio Natriegens. ACS Synth. Biol. 2021, 10 (7), 1615–1624. 10.1021/acssynbio.1c00110. [DOI] [PubMed] [Google Scholar]
- Yim S. S.; Johns N. I.; Noireaux V.; Wang H. H. Protecting Linear DNA Templates in Cell-Free Expression Systems from Diverse Bacteria. ACS Synth. Biol. 2020, 9, 2851. 10.1021/acssynbio.0c00277. [DOI] [PubMed] [Google Scholar]
- Wu P. S. C.; Ozawa K.; Lim S. P.; Vasudevan S. G.; Dixon N. E.; Otting G. Cell-Free Transcription/Translation from PCR-Amplified DNA for High-Throughput NMR Studies. Angew. Chem., Int. Ed. Engl. 2007, 46 (18), 3356–3358. 10.1002/anie.200605237. [DOI] [PubMed] [Google Scholar]
- Cole S. D.; Beabout K.; Turner K. B.; Smith Z. K.; Funk V. L.; Harbaugh S. V.; Liem A. T.; Roth P. A.; Geier B. A.; Emanuel P. A.; Walper S. A.; Chávez J. L.; Lux M. W. Quantification of Interlaboratory Cell-Free Protein Synthesis Variability. ACS Synth. Biol. 2019, 8 (9), 2080–2091. 10.1021/acssynbio.9b00178. [DOI] [PubMed] [Google Scholar]
- Romantseva E.; Alperovich N.; Ross D.; Lund S. P.; Strychalski E. A. Effects of DNA Template Preparation on Variability in Cell-Free Protein Production. Synth. Biol. 2022, 7 (1), ysac015. 10.1093/synbio/ysac015. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Romantseva E. F.; Tack D. S.; Alperovich N.; Ross D.; Strychalski E. A.. Best Practices for toward Improved Reproducibility In. In Cell-Free Gene Expression; Methods in Molecular Biology (Clifton, N.J.); Springer US: New York, NY, 2022; pp 3–50. 10.1007/978-1-0716-1998-8_1. [DOI] [PubMed] [Google Scholar]
- Chappell J.; Jensen K.; Freemont P. S. Validation of an Entirely in Vitro Approach for Rapid Prototyping of DNA Regulatory Elements for Synthetic Biology. Nucleic Acids Res. 2013, 41 (5), 3471–3481. 10.1093/nar/gkt052. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McManus J. B.; Emanuel P. A.; Murray R. M.; Lux M. W. A Method for Cost-Effective and Rapid Characterization of Engineered T7-Based Transcription Factors by Cell-Free Protein Synthesis Reveals Insights into the Regulation of T7 RNA Polymerase-Driven Expression. Arch. Biochem. Biophys. 2019, 674, 108045. 10.1016/j.abb.2019.07.010. [DOI] [PubMed] [Google Scholar]
- Su’etsugu M.; Takada H.; Katayama T.; Tsujimoto H. Exponential Propagation of Large Circular DNA by Reconstitution of a Chromosome-Replication Cycle. Nucleic Acids Res. 2017, 45 (20), 11525–11534. 10.1093/nar/gkx822. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Grasemann L.; Thiel Pizarro P.; Maerkl S. J. C2CAplus: A One-Pot Isothermal Circle-to-Circle DNA Amplification System. ACS Synth. Biol. 2023, 12 (10), 3137–3142. 10.1021/acssynbio.3c00390. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kung H. F.; Redfield B.; Treadwell B. V.; Eskin B.; Spears C.; Weissbach H. DNA-Directed in Vitro Synthesis of Beta-Galactosidase. Studies with Purified Factors. J. Biol. Chem. 1977, 252 (19), 6889–6894. 10.1016/S0021-9258(17)39933-7. [DOI] [PubMed] [Google Scholar]
- Ganoza M. C.; Cunningham C.; Green R. M. Isolation and Point of Action of a Factor from Escherichia Coli Required to Reconstruct Translation. Proc. Natl. Acad. Sci. U. S. A. 1985, 82 (6), 1648–1652. 10.1073/pnas.82.6.1648. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pavlov M. Y.; Ehrenberg M. Rate of Translation of Natural MRNAs in an Optimized in Vitro System. Arch. Biochem. Biophys. 1996, 328 (1), 9–16. 10.1006/abbi.1996.0136. [DOI] [PubMed] [Google Scholar]
- Jewett M. C.; Forster A. C. Update on Designing and Building Minimal Cells. Curr. Opin. Biotechnol. 2010, 21 (5), 697–703. 10.1016/j.copbio.2010.06.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kuruma Y.; Ueda T. The PURE System for the Cell-Free Synthesis of Membrane Proteins. Nat. Protoc. 2015, 10 (9), 1328–1344. 10.1038/nprot.2015.082. [DOI] [PubMed] [Google Scholar]
- Maddalena L. L. de; Niederholtmeyer H.; Turtola M.; Swank Z. N.; Belogurov G. A.; Maerkl S. J. GreA and GreB Enhance Expression of Escherichia Coli RNA Polymerase Promoters in a Reconstituted Transcription-Translation System. ACS Synth. Biol. 2016, 5 (9), 929–935. 10.1021/acssynbio.6b00017. [DOI] [PubMed] [Google Scholar]
- PUREfrex®2.0. genefrontier. https://purefrex.genefrontier.com/products/cellfreeproteinsynthesiskits/purefrex2.0.html (accessed 2023-07-05).
- New England Biolabs . PURExpress® in vitro protein synthesis kit. https://www.neb.com/products/e6800-purexpress-invitro-protein-synthesis-kit (accessed 2023-07-05).
- MyTXTL - cell-free protein expression. Daicel Arbor Biosciences. https://arborbiosci.com/synthetic-biology/cell-free-protein-expression/mytxtl/ (accessed 2023-07-05).
- New England Biolabs . NEBExpress Cell-free ecoli Protein Synthesis System. https://www.neb.com/products/e5360-nebexpress-cell-free-ecoli-protein-synthesis-system (accessed 2023-07-05).
- Villarreal F.; Contreras-Llano L. E.; Chavez M.; Ding Y.; Fan J.; Pan T.; Tan C. Synthetic Microbial Consortia Enable Rapid Assembly of Pure Translation Machinery. Nat. Chem. Biol. 2018, 14 (1), 29–35. 10.1038/nchembio.2514. [DOI] [PubMed] [Google Scholar]
- Lavickova B.; Maerkl S. J. A Simple, Robust, and Low-Cost Method To Produce the PURE Cell-Free System. ACS Synth. Biol. 2019, 8 (2), 455–462. 10.1021/acssynbio.8b00427. [DOI] [PubMed] [Google Scholar]
- Lee K.-H.; Kwon Y.-C.; Yoo S. J.; Kim D.-M. Ribosomal Synthesis and in Situ Isolation of Peptide Molecules in a Cell-Free Translation System. Protein Expr. Purif. 2010, 71 (1), 16–20. 10.1016/j.pep.2010.01.016. [DOI] [PubMed] [Google Scholar]
- Wu Y.; Cui Z.; Huang Y.-H.; de Veer S. J.; Aralov A. V.; Guo Z.; Moradi S. V.; Hinton A. O.; Deuis J. R.; Guo S.; Chen K.-E.; Collins B. M.; Vetter I.; Herzig V.; Jones A.; Cooper M. A.; King G. F.; Craik D. J.; Alexandrov K.; Mureev S.. Towards a Generic Prototyping Approach for Therapeutically-Relevant Peptides and Proteins in a Cell-Free Translation System. Nat. Commun. 2022, 13 ( (1), ). 10.1038/s41467-021-27854-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- DeWinter M. A.; Thames A. H.; Guerrero L.; Kightlinger W.; Karim A. S.; Jewett M. C. Point-of-Care Peptide Hormone Production Enabled by Cell-Free Protein Synthesis. ACS Synth. Biol. 2023, 12 (4), 1216–1226. 10.1021/acssynbio.2c00680. [DOI] [PubMed] [Google Scholar]
- Huang Y.; Wiedmann M. M.; Suga H. RNA Display Methods for the Discovery of Bioactive Macrocycles. Chem. Rev. 2019, 119 (17), 10360–10391. 10.1021/acs.chemrev.8b00430. [DOI] [PubMed] [Google Scholar]
- Wang H. H.; Huang P. Y.; Xu G.; Haas W.; Marblestone A.; Li J.; Gygi S. P.; Forster A. C.; Jewett M. C.; Church G. M. Multiplexed in Vivo His-Tagging of Enzyme Pathways for in Vitro Single-Pot Multienzyme Catalysis. ACS Synth. Biol. 2012, 1 (2), 43–52. 10.1021/sb3000029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shepherd T. R.; Du L.; Liljeruhm J.; Samudyata; Wang J.; Sjödin M. O. D.; Wetterhall M.; Yomo T.; Forster A. C. De Novodesign and Synthesis of a 30-Cistron Translation-Factor Module. Nucleic Acids Res. 2017, 45 (18), 10895–10905. 10.1093/nar/gkx753. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ngo P. H. T.; Ishida S.; Busogi B. B.; Do H.; Ledesma M. A.; Kar S.; Ellington A. Changes in Coding and Efficiency through Modular Modifications to a One Pot PURE System for in Vitro Transcription and Translation. ACS Synth. Biol. 2023, 12 (12), 3771–3777. 10.1021/acssynbio.3c00461. [DOI] [PubMed] [Google Scholar]
- Li J.; Gu L.; Aach J.; Church G. M.. Improved Cell-Free RNA and Protein Synthesis System. PLoS One 2014, 9 ( (9), ). 10.1371/journal.pone.0106232. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kazuta Y.; Matsuura T.; Ichihashi N.; Yomo T. Synthesis of Milligram Quantities of Proteins Using a Reconstituted in Vitro Protein Synthesis System. J. Biosci. Bioeng. 2014, 118 (5), 554–557. 10.1016/j.jbiosc.2014.04.019. [DOI] [PubMed] [Google Scholar]
- Matsuura T.; Kazuta Y.; Aita T.; Adachi J.; Yomo T. Quantifying Epistatic Interactions among the Components Constituting the Protein Translation System. Mol. Syst. Biol. 2009, 5 (1), 297. 10.1038/msb.2009.50. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Moriizumi Y.; Tabata K. V.; Miyoshi D.; Noji H. Osmolyte-Enhanced Protein Synthesis Activity of a Reconstituted Translation System. ACS Synth. Biol. 2019, 8 (3), 557–567. 10.1021/acssynbio.8b00513. [DOI] [PubMed] [Google Scholar]
- Ganesh R. B.; Maerkl S. J. Towards Self-Regeneration: Exploring the Limits of Protein Synthesis in the PURE Cell-Free Transcription-Translation System. bioRxiv 2024, 13, 2555. 10.1021/acssynbio.4c00304. [DOI] [PubMed] [Google Scholar]
- Sheahan T.; Wieden H.-J. Ribosomal Protein S1 Improves the Protein Yield of an in Vitro Reconstituted Cell-Free Translation System. ACS Synth. Biol. 2022, 11 (2), 1004–1008. 10.1021/acssynbio.1c00514. [DOI] [PubMed] [Google Scholar]
- Ying B.-W.; Taguchi H.; Kondo M.; Ueda T. Co-Translational Involvement of the Chaperonin GroEL in the Folding of Newly Translated Polypeptides. J. Biol. Chem. 2005, 280 (12), 12035–12040. 10.1074/jbc.M500364200. [DOI] [PubMed] [Google Scholar]
- Wang P.-H.; Fujishima K.; Berhanu S.; Kuruma Y.; Jia T. Z.; Khusnutdinova A. N.; Yakunin A. F.; McGlynn S. E. A Bifunctional Polyphosphate Kinase Driving the Regeneration of Nucleoside Triphosphate and Reconstituted Cell-Free Protein Synthesis. ACS Synth. Biol. 2020, 9 (1), 36–42. 10.1021/acssynbio.9b00456. [DOI] [PubMed] [Google Scholar]
- Damiati S.; Mhanna R.; Kodzius R.; Ehmoser E.-K. Cell-Free Approaches in Synthetic Biology Utilizing Microfluidics. Genes (Basel) 2018, 9 (3), 144. 10.3390/genes9030144. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Endo Y.; Otsuzuki S.; Ito K.; Miura K.-I. Production of an Enzymatic Active Protein Using a Continuous Flow Cell-Free Translation System. J. Biotechnol. 1992, 25 (3), 221–230. 10.1016/0168-1656(92)90157-5. [DOI] [PubMed] [Google Scholar]
- van Sluijs B.; Maas R. J. M.; van der Linden A. J.; de Greef T. F. A.; Huck W. T. S.. A Microfluidic Optimal Experimental Design Platform for Forward Design of Cell-Free Genetic Networks. Nat. Commun. 2022, 13 ( (1), ). 10.1038/s41467-022-31306-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Karzbrun E.; Tayar A. M.; Noireaux V.; Bar-Ziv R. H. Programmable On-Chip DNA Compartments as Artificial Cells. Science 2014, 345 (6198), 829–832. 10.1126/science.1255550. [DOI] [PubMed] [Google Scholar]
- Yang D.; Peng S.; Hartman M. R.; Gupton-Campolongo T.; Rice E. J.; Chang A. K.; Gu Z.; Lu G. Q.; Luo D.. Enhanced Transcription and Translation in Clay Hydrogel and Implications for Early Life Evolution. Sci. Rep. 2013, 3 ( (1), ). 10.1038/srep03165. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Niederholtmeyer H.; Chaggan C.; Devaraj N. K. Communication and Quorum Sensing in Non-Living Mimics of Eukaryotic Cells. Nat. Commun. 2018, 9 (1), 5027. 10.1038/s41467-018-07473-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Blum S. M.; Lee M. S.; Mgboji G. E.; Funk V. L.; Beabout K.; Harbaugh S. V.; Roth P. A.; Liem A. T.; Miklos A. E.; Emanuel P. A.; Walper S. A.; Chávez J. L.; Lux M. W. Impact of Porous Matrices and Concentration by Lyophilization on Cell-Free Expression. ACS Synth. Biol. 2021, 10, 1116. 10.1021/acssynbio.0c00634. [DOI] [PubMed] [Google Scholar]
- Pardee K.; Green A. A.; Ferrante T.; Cameron D. E.; DaleyKeyser A.; Yin P.; Collins J. J. Paper-Based Synthetic Gene Networks. Cell 2014, 159 (4), 940–954. 10.1016/j.cell.2014.10.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Pardee K.; Green A. A.; Takahashi M. K.; Braff D.; Lambert G.; Lee J. W.; Ferrante T.; Ma D.; Donghia N.; Fan M.; Daringer N. M.; Bosch I.; Dudley D. M.; O’Connor D. H.; Gehrke L.; Collins J. J. Rapid, Low-Cost Detection of Zika Virus Using Programmable Biomolecular Components. Cell 2016, 165 (5), 1255–1266. 10.1016/j.cell.2016.04.059. [DOI] [PubMed] [Google Scholar]
- Whitfield C. J.; Banks A. M.; Dura G.; Love J.; Fieldsend J. E.; Goodchild S. A.; Fulton D. A.; Howard T. P. Cell-Free Protein Synthesis in Hydrogel Materials. Chem. Commun. (Camb.) 2020, 56 (52), 7108–7111. 10.1039/D0CC02582H. [DOI] [PubMed] [Google Scholar]
- Benítez-Mateos A. I.; Zeballos N.; Comino N.; Moreno de Redrojo L.; Randelovic T.; López-Gallego F. Microcompartmentalized Cell-Free Protein Synthesis in Hydrogel μ-Channels. ACS Synth. Biol. 2020, 9 (11), 2971–2978. 10.1021/acssynbio.0c00462. [DOI] [PubMed] [Google Scholar]
- Cui J.; Wu D.; Sun Q.; Yang X.; Wang D.; Zhuang M.; Zhang Y.; Gan M.; Luo D.. A PEGDA/DNA Hybrid Hydrogel for Cell-Free Protein Synthesis. Front. Chem. 2020, 8. 10.3389/fchem.2020.00028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heida T.; Köhler T.; Kaufmann A.; Männel M. J.; Thiele J.. Cell-free Protein Synthesis in Bifunctional Hyaluronan Microgels: A Strategy for in Situ Immobilization and Purification of His-tagged Proteins. ChemSystemsChem. 2020, 2 ( (3), ). 10.1002/syst.201900058. [DOI] [Google Scholar]
- Lee M. S.; Lee J. A.; Biondo J. R.; Lux J. E.; Raig R. M.; Berger P. N.; Bernhards C. B.; Kuhn D. L.; Gupta M. K.; Lux M. W. Cell-Free Protein Expression in Polymer Materials. ACS Synth. Biol. 2024, 13 (4), 1152–1164. 10.1021/acssynbio.3c00628. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ho G.; Kubušová V.; Irabien C.; Li V.; Weinstein A.; Chawla S.; Yeung D.; Mershin A.; Zolotovsky K.; Mogas-Soldevila L.. Multiscale Design of Cell-Free Biologically Active Architectural Structures. Front. Bioeng. Biotechnol. 2023, 11. 10.3389/fbioe.2023.1125156. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nguyen P. Q.; Soenksen L. R.; Donghia N. M.; Angenent-Mari N. M.; de Puig H.; Huang A.; Lee R.; Slomovic S.; Galbersanini T.; Lansberry G.; Sallum H. M.; Zhao E. M.; Niemi J. B.; Collins J. J. Wearable Materials with Embedded Synthetic Biology Sensors for Biomolecule Detection. Nat. Biotechnol. 2021, 39 (11), 1366–1374. 10.1038/s41587-021-00950-3. [DOI] [PubMed] [Google Scholar]
- Kelwick R. J. R.; Webb A. J.; Freemont P. S.. Biological Materials: The next Frontier for Cell-Free Synthetic Biology. Front. Bioeng. Biotechnol. 2020, 8. 10.3389/fbioe.2020.00399. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Choi Y.-N.; Cho N.; Lee K.; Gwon D.-A.; Lee J. W.; Lee J.. Programmable Synthesis of Biobased Materials Using Cell-free Systems. Adv. Mater. 2023, 35 ( (4), ). 10.1002/adma.202203433. [DOI] [PubMed] [Google Scholar]
- Swartz J. R.; Jewett M. C.; Woodrow K. A.. Cell-Free Protein Synthesis With Prokaryotic Combined Transcription-Translation. In Recombinant Gene Expression: Reviews and Protocols; Balbás P., Lorence A., Eds.; Humana Press: Totowa, NJ, 2004; pp 169–182. 10.1385/1-59259-774-2:169. [DOI] [PubMed] [Google Scholar]
- Siegal-Gaskins D.; Tuza Z. A.; Kim J.; Noireaux V.; Murray R. M. Gene Circuit Performance Characterization and Resource Usage in a Cell-Free “Breadboard.. ACS Synth. Biol. 2014, 3 (6), 416–425. 10.1021/sb400203p. [DOI] [PubMed] [Google Scholar]
- Jung J. K.; Alam K. K.; Verosloff M. S.; Capdevila D. A.; Desmau M.; Clauer P. R.; Lee J. W.; Nguyen P. Q.; Pastén P. A.; Matiasek S. J.; Gaillard J.-F.; Giedroc D. P.; Collins J. J.; Lucks J. B. Cell-Free Biosensors for Rapid Detection of Water Contaminants. Nat. Biotechnol. 2020, 38 (12), 1451–1459. 10.1038/s41587-020-0571-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen Z.; Kibler R. D.; Hunt A.; Busch F.; Pearl J.; Jia M.; VanAernum Z. L.; Wicky B. I. M.; Dods G.; Liao H.; Wilken M. S.; Ciarlo C.; Green S.; El-Samad H.; Stamatoyannopoulos J.; Wysocki V. H.; Jewett M. C.; Boyken S. E.; Baker D. De Novo Design of Protein Logic Gates. Science 2020, 368 (6486), 78–84. 10.1126/science.aay2790. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martemyanov K. A.; Shirokov V. A.; Kurnasov O. V.; Gudkov A. T.; Spirin A. S. Cell-Free Production of Biologically Active Polypeptides: Application to the Synthesis of Antibacterial Peptide Cecropin. Protein Expr. Purif. 2001, 21 (3), 456–461. 10.1006/prep.2001.1400. [DOI] [PubMed] [Google Scholar]
- Yuan Q.; Wu M.; Liao Y.; Liang S.; Lu Y.; Lin Y. Rapid Prototyping Enzyme Homologs to Improve Titer of Nicotinamide Mononucleotide Using a Strategy Combining Cell-free Protein Synthesis with Split GFP. Biotechnol. Bioeng. 2023, 120 (4), 1133–1146. 10.1002/bit.28326. [DOI] [PubMed] [Google Scholar]
- Wick S.; Walsh D. I. 3rd; Bobrow J.; Hamad-Schifferli K.; Kong D. S.; Thorsen T.; Mroszczyk K.; Carr P. A. PERSIA for Direct Fluorescence Measurements of Transcription, Translation, and Enzyme Activity in Cell-Free Systems. ACS Synth. Biol. 2019, 8 (5), 1010–1025. 10.1021/acssynbio.8b00450. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wick S.; Carr P. A. Measurement of Transcription, Translation, and Other Enzymatic Processes during Cell-Free Expression Using PERSIA. Methods Mol. Biol. 2022, 2433, 169–181. 10.1007/978-1-0716-1998-8_10. [DOI] [PubMed] [Google Scholar]
- Willi J. A.; Jewett M. C. Fluorescent Minihelix Assay for Translation Quantification in High-Throughput Cell-Free Systems. bioRxiv 2023, 10.1101/2023.10.13.562088. [DOI] [Google Scholar]
- Jaroentomeechai T.; Stark J. C.; Natarajan A.; Glasscock C. J.; Yates L. E.; Hsu K. J.; Mrksich M.; Jewett M. C.; DeLisa M. P.. Single-Pot Glycoprotein Biosynthesis Using a Cell-Free Transcription-Translation System Enriched with Glycosylation Machinery. Nat. Commun. 2018, 9 ( (1), ). 10.1038/s41467-018-05110-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schoborg J. A.; Hershewe J. M.; Stark J. C.; Kightlinger W.; Kath J. E.; Jaroentomeechai T.; Natarajan A.; DeLisa M. P.; Jewett M. C. A Cell-Free Platform for Rapid Synthesis and Testing of Active Oligosaccharyltransferases. Biotechnol. Bioeng. 2018, 115 (3), 739–750. 10.1002/bit.26502. [DOI] [PubMed] [Google Scholar]
- Rasor B. J.; Vögeli B.; Jewett M. C.; Karim A. S.. Cell-Free Protein Synthesis for High-Throughput Biosynthetic Prototyping. In Cell-Free Gene Expression; Methods in Molecular Biology (Clifton, N.J.); Springer US: New York, NY, 2022; pp 199–215. 10.1007/978-1-0716-1998-8_12. [DOI] [PubMed] [Google Scholar]
- Cabantous S.; Waldo G. S. In Vivo and in Vitro Protein Solubility Assays Using Split GFP. Nat. Methods 2006, 3 (10), 845–854. 10.1038/nmeth932. [DOI] [PubMed] [Google Scholar]
- Liu H.; Naismith J. H. A Simple and Efficient Expression and Purification System Using Two Newly Constructed Vectors. Protein Expr. Purif. 2009, 63 (2), 102–111. 10.1016/j.pep.2008.09.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boute N.; Lowe P.; Berger S.; Malissard M.; Robert A.; Tesar M. NanoLuc Luciferase - A Multifunctional Tool for High Throughput Antibody Screening. Front. Pharmacol. 2016, 7 (FEB), 1–11. 10.3389/fphar.2016.00027. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Niederholtmeyer H.; Xu L.; Maerkl S. J. Real-Time MRNA Measurement during an in Vitro Transcription and Translation Reaction Using Binary Probes. ACS Synth. Biol. 2013, 2 (8), 411–417. 10.1021/sb300104f. [DOI] [PubMed] [Google Scholar]
- Hook B.; Schagat T. Non-Radioactive Detection of Proteins Expressed in Cell-Free Expression Systems. https://www.promega.com/resources/pubhub/tpub_049-nonradioactive-detection-of-proteins-expressed-in-cell-free-expression-systems (accessed 2022-03-28).
- Kightlinger W.; Duncker K. E.; Ramesh A.; Thames A. H.; Natarajan A.; Stark J. C.; Yang A.; Lin L.; Mrksich M.; DeLisa M. P.; Jewett M. C. A Cell-Free Biosynthesis Platform for Modular Construction of Protein Glycosylation Pathways. Nat. Commun. 2019, 10 (1), 5404. 10.1038/s41467-019-12024-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vinarov D. A.; Newman C. L. L.; Markley J. L. Wheat Germ Cell-Free Platform for Eukaryotic Protein Production. FEBS J. 2006, 273 (18), 4160–4169. 10.1111/j.1742-4658.2006.05434.x. [DOI] [PubMed] [Google Scholar]
- Makino S.-I.; Goren M. A.; Fox B. G.; Markley J. L. Cell-Free Protein Synthesis Technology in NMR High-Throughput Structure Determination. Methods Mol. Biol. 2010, 607, 127–147. 10.1007/978-1-60327-331-2_12. [DOI] [PubMed] [Google Scholar]
- Aoki M.; Matsuda T.; Tomo Y.; Miyata Y.; Inoue M.; Kigawa T.; Yokoyama S. Automated System for High-Throughput Protein Production Using the Dialysis Cell-Free Method. Protein Expr. Purif. 2009, 68 (2), 128–136. 10.1016/j.pep.2009.07.017. [DOI] [PubMed] [Google Scholar]
- Hughes R. A.; Ellington A. D. Synthetic DNA Synthesis and Assembly: Putting the Synthetic in Synthetic Biology. Cold Spring Harb. Perspect. Biol. 2017, 9 (1), a023812. 10.1101/cshperspect.a023812. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hunt A. C.; Case J. B.; Park Y.-J.; Cao L.; Wu K.; Walls A. C.; Liu Z.; Bowen J. E.; Yeh H.-W.; Saini S.; Helms L.; Zhao Y. T.; Hsiang T.-Y.; Starr T. N.; Goreshnik I.; Kozodoy L.; Carter L.; Ravichandran R.; Green L. B.; Matochko W. L.; Thomson C. A.; Vögeli B.; Krüger A.; VanBlargan L. A.; Chen R. E.; Ying B.; Bailey A. L.; Kafai N. M.; Boyken S. E.; Ljubetič A.; Edman N.; Ueda G.; Chow C. M.; Johnson M.; Addetia A.; Navarro M. J.; Panpradist N.; Gale M. Jr; Freedman B. S.; Bloom J. D.; Ruohola-Baker H.; Whelan S. P. J.; Stewart L.; Diamond M. S.; Veesler D.; Jewett M. C.; Baker D. Multivalent Designed Proteins Neutralize SARS-CoV-2 Variants of Concern and Confer Protection against Infection in Mice. Sci. Transl. Med. 2022, 14, eabn1252 10.1126/scitranslmed.abn1252. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Caschera F.; Karim A. S.; Gazzola G.; D’Aquino A. E.; Packard N. H.; Jewett M. C. High-Throughput Optimization Cycle of a Cell-Free Ribosome Assembly and Protein Synthesis System. ACS Synth. Biol. 2018, 7 (12), 2841–2853. 10.1021/acssynbio.8b00276. [DOI] [PubMed] [Google Scholar]
- Karim A. S.; Heggestad J. T.; Crowe S. A.; Jewett M. C. Controlling Cell-Free Metabolism through Physiochemical Perturbations. Metab. Eng. 2018, 45, 86–94. 10.1016/j.ymben.2017.11.005. [DOI] [PubMed] [Google Scholar]
- Caschera F.; Bedau M. A.; Buchanan A.; Cawse J.; de Lucrezia D.; Gazzola G.; Hanczyc M. M.; Packard N. H. Coping with Complexity: Machine Learning Optimization of Cell-Free Protein Synthesis. Biotechnol. Bioeng. 2011, 108 (9), 2218–2228. 10.1002/bit.23178. [DOI] [PubMed] [Google Scholar]
- Rapp J. T.; Bremer B. J.; Romero P. A. Self-Driving Laboratories to Autonomously Navigate the Protein Fitness Landscape. Nat. Chem. Eng. 2024, 1 (1), 97–107. 10.1038/s44286-023-00002-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Marshall R.; Garamella J.; Noireaux V.; Pierson A.. High-Throughput Microliter-Sized Cell-Free Transcription-Translation Reactions for Synthetic Biology Applications Using the Echo® 550 Liquid Handler. Labcyte Appl. Note 2018. [Google Scholar]
- Borkowski O.; Koch M.; Zettor A.; Pandi A.; Batista A. C.; Soudier P.; Faulon J.-L. Large Scale Active-Learning-Guided Exploration for in Vitro Protein Production Optimization. Nat. Commun. 2020, 11 (1), 1872. 10.1038/s41467-020-15798-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hammerling M. J.; Fritz B. R.; Yoesep D. J.; Kim D. S.; Carlson E. D.; Jewett M. C. In Vitro Ribosome Synthesis and Evolution through Ribosome Display. Nat. Commun. 2020, 11 (1), 1108. 10.1038/s41467-020-14705-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Marshall R.; Noireaux V. Quantitative Modeling of Transcription and Translation of an All-E. Coli Cell-Free System. Sci. Rep. 2019, 9 (1), 11980. 10.1038/s41598-019-48468-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Angenendt P.; Nyarsik L.; Szaflarski W.; Glökler J.; Nierhaus K. H.; Lehrach H.; Cahill D. J.; Lueking A. Cell-Free Protein Expression and Functional Assay in Nanowell Chip Format. Anal. Chem. 2004, 76 (7), 1844–1849. 10.1021/ac035114i. [DOI] [PubMed] [Google Scholar]
- He M. Single Step Generation of Protein Arrays from DNA by Cell-Free Expression and in Situ Immobilisation (PISA Method). Nucleic Acids Res. 2001, 29 (15), 73e–773. 10.1093/nar/29.15.e73. [DOI] [PMC free article] [PubMed] [Google Scholar]
- He M.; Taussig M. J. DiscernArray Technology: A Cell-Free Method for the Generation of Protein Arrays from PCR DNA. J. Immunol. Methods 2003, 274 (1–2), 265–270. 10.1016/S0022-1759(02)00521-5. [DOI] [PubMed] [Google Scholar]
- Angenendt P.; Kreutzberger J.; Glökler J.; Hoheisel J. D. Generation of High Density Protein Microarrays by Cell-Free in Situ Expression of Unpurified PCR Products. Mol. Cell. Proteomics 2006, 5 (9), 1658–1666. 10.1074/mcp.T600024-MCP200. [DOI] [PubMed] [Google Scholar]
- Ramachandran N. Self-Assembling Protein Microarrays. Science 2004, 305 (5680), 86–90. 10.1126/science.1097639. [DOI] [PubMed] [Google Scholar]
- Ramachandran N.; Raphael J. V.; Hainsworth E.; Demirkan G.; Fuentes M. G.; Rolfs A.; Hu Y.; LaBaer J. Next-Generation High-Density Self-Assembling Functional Protein Arrays. Nat. Methods 2008, 5 (6), 535–538. 10.1038/nmeth.1210. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zárate X.; Henderson D. C.; Phillips K. C.; Lake A. D.; Galbraith D. W. Development of High-Yield Autofluorescent Protein Microarrays Using Hybrid Cell-Free Expression with Combined Escherichia Coli S30 and Wheat Germ Extracts. Proteome Sci. 2010, 8 (1), 32. 10.1186/1477-5956-8-32. [DOI] [PMC free article] [PubMed] [Google Scholar]
- He M.; Stoevesandt O.; Palmer E. A.; Khan F.; Ericsson O.; Taussig M. J. Printing Protein Arrays from DNA Arrays. Nat. Methods 2008, 5 (2), 175–177. 10.1038/nmeth.1178. [DOI] [PubMed] [Google Scholar]
- Biyani M.; Osawa T.; Nemoto N.; Ichiki T. Microintaglio Printing of Biomolecules and Its Application ToIn SituProduction of Messenger Ribonucleic Acid Display Microarray. Appl. Phys. Express 2011, 4 (4), 047001. 10.1143/APEX.4.047001. [DOI] [Google Scholar]
- Biyani M.; Moriyasu J.; Tanaka Y.; Sato S.; Ueno S.; Ichiki T. Microintaglio Printing of in Situ Synthesized Proteins Enables Rapid Printing of High-Density Protein Microarrays Directly from DNA Microarrays. Appl. Phys. Express 2013, 6 (8), 087001. 10.7567/APEX.6.087001. [DOI] [Google Scholar]
- He M.; Stoevesandt O.; Taussig M. J. In Situ Synthesis of Protein Arrays. Curr. Opin. Biotechnol. 2008, 19 (1), 4–9. 10.1016/j.copbio.2007.11.009. [DOI] [PubMed] [Google Scholar]
- Contreras-Llano L. E.; Tan C. High-Throughput Screening of Biomolecules Using Cell-Free Gene Expression Systems. Synth. Biol. 2018, 3 (1), 1–13. 10.1093/synbio/ysy012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Swank Z.; Laohakunakorn N.; Maerkl S. J. Cell-Free Gene-Regulatory Network Engineering with Synthetic Transcription Factors. Proc. Natl. Acad. Sci. U. S. A. 2019, 116 (13), 5892–5901. 10.1073/pnas.1816591116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nand A.; Singh V.; Pérez J. B.; Tyagi D.; Cheng Z.; Zhu J. In Situ Protein Microarrays Capable of Real-Time Kinetics Analysis Based on Surface Plasmon Resonance Imaging. Anal. Biochem. 2014, 464, 30–35. 10.1016/j.ab.2014.06.002. [DOI] [PubMed] [Google Scholar]
- Maerkl S. J.; Quake S. R. A Systems Approach to Measuring the Binding Energy Landscapes of Transcription Factors. Science 2007, 315 (5809), 233–237. 10.1126/science.1131007. [DOI] [PubMed] [Google Scholar]
- Maerkl S. J.; Quake S. R. Experimental Determination of the Evolvability of a Transcription Factor. Proc. Natl. Acad. Sci. U. S. A. 2009, 106 (44), 18650–18655. 10.1073/pnas.0907688106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gerber D.; Maerkl S. J.; Quake S. R. An in Vitro Microfluidic Approach to Generating Protein-Interaction Networks. Nat. Methods 2009, 6 (1), 71–74. 10.1038/nmeth.1289. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Heyman Y.; Buxboim A.; Wolf S. G.; Daube S. S.; Bar-Ziv R. H. Cell-Free Protein Synthesis and Assembly on a Biochip. Nat. Nanotechnol. 2012, 7 (6), 374–378. 10.1038/nnano.2012.65. [DOI] [PubMed] [Google Scholar]
- Levy M.; Falkovich R.; Daube S. S.; Bar-Ziv R. H. Autonomous Synthesis and Assembly of a Ribosomal Subunit on a Chip. Sci. Adv. 2020, 6 (16), eaaz6020 10.1126/sciadv.aaz6020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Markin C. J.; Mokhtari D. A.; Sunden F.; Appel M. J.; Akiva E.; Longwell S. A.; Sabatti C.; Herschlag D.; Fordyce P. M. Revealing Enzyme Functional Architecture via High-Throughput Microfluidic Enzyme Kinetics. Science 2021, 373 (6553), eabf8761 10.1126/science.abf8761. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Greiss F.; Lardon N.; Schütz L.; Barak Y.; Daube S. S.; Weinhold E.; Noireaux V.; Bar-Ziv R. A Genetic Circuit on a Single DNA Molecule as an Autonomous Dissipative Nanodevice. Nat. Commun. 2024, 15 (1), 1–12. 10.1038/s41467-024-45186-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tayar A. M.; Karzbrun E.; Noireaux V.; Bar-Ziv R. H. Synchrony and Pattern Formation of Coupled Genetic Oscillators on a Chip of Artificial Cells. Proc. Natl. Acad. Sci. U. S. A. 2017, 114 (44), 11609–11614. 10.1073/pnas.1710620114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Greiss F.; Daube S. S.; Noireaux V.; Bar-Ziv R. From Deterministic to Fuzzy Decision-Making in Artificial Cells. Nat. Commun. 2020, 11 (1), 1–9. 10.1038/s41467-020-19395-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tayar A. M.; Karzbrun E.; Noireaux V.; Bar-Ziv R. H. Propagating Gene Expression Fronts in a One-Dimensional Coupled System of Artificial Cells. Nat. Phys. 2015, 11 (12), 1037–1041. 10.1038/nphys3469. [DOI] [Google Scholar]
- Tayar A. M.; Daube S. S.; Bar-Ziv R. H. Progress in Programming Spatiotemporal Patterns and Machine-Assembly in Cell-Free Protein Expression Systems. Curr. Opin. Chem. Biol. 2017, 40, 37–46. 10.1016/j.cbpa.2017.05.005. [DOI] [PubMed] [Google Scholar]
- Vonshak O.; Divon Y.; Förste S.; Garenne D.; Noireaux V.; Lipowsky R.; Rudorf S.; Daube S. S.; Bar-Ziv R. H. Programming Multi-Protein Assembly by Gene-Brush Patterns and Two-Dimensional Compartment Geometry. Nat. Nanotechnol. 2020, 15, 783. 10.1038/s41565-020-0720-7. [DOI] [PubMed] [Google Scholar]
- Levy M.; Falkovich R.; Vonshak O.; Bracha D.; Tayar A. M.; Shimizu Y.; Daube S. S.; Bar-Ziv R. H. Boundary-Free Ribosome Compartmentalization by Gene Expression on a Surface. ACS Synth. Biol. 2021, 10 (3), 609–619. 10.1021/acssynbio.0c00613. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang Y.; Minagawa Y.; Kizoe H.; Miyazaki K.; Iino R.; Ueno H.; Tabata K. V.; Shimane Y.; Noji H.. Accurate High-Throughput Screening Based on Digital Protein Synthesis in a Massively Parallel Femtoliter Droplet Array. Sci. Adv. 2019, No. 5. 10.1126/sciadv.aav8185. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Layton C. J.; McMahon P. L.; Greenleaf W. J. Large-Scale, Quantitative Protein Assays on a High-Throughput DNA Sequencing Chip. Mol. Cell 2019, 73 (5), 1075–1082. 10.1016/j.molcel.2019.02.019. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Porebski B. T.; Balmforth M.; Browne G.; Riley A.; Jamali K.; Fürst M. J. L. J.; Velic M.; Buchanan A.; Minter R.; Vaughan T.; Holliger P. Rapid Discovery of High-Affinity Antibodies via Massively Parallel Sequencing, Ribosome Display and Affinity Screening. Nat. Biomed. Eng. 2024, 8, 214. 10.1038/s41551-023-01093-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Baranwal A. K.; Maerkl S. J. A Comprehensive Review of Microfluidic Approaches in Cell-Free Synthetic Biology. Front. Synth. Biol. 2024, 2, 1397533. 10.3389/fsybi.2024.1397533. [DOI] [Google Scholar]
- Courtois F.; Olguin L. F.; Whyte G.; Bratton D.; Huck W. T. S.; Abell C.; Hollfelder F. An Integrated Device for Monitoring Time-Dependent in Vitro Expression from Single Genes in Picolitre Droplets. Chembiochem 2008, 9 (3), 439–446. 10.1002/cbic.200700536. [DOI] [PubMed] [Google Scholar]
- Tawfik D. S.; Griffiths A. D. Man-Made Cell-like Compartments for Molecular Evolution. Nat. Biotechnol. 1998, 16 (7), 652–656. 10.1038/nbt0798-652. [DOI] [PubMed] [Google Scholar]
- Griffiths A. D.; Tawfik D. S. Directed Evolution of an Extremely Fast Phosphotriesterase by in Vitro Compartmentalization. EMBO J. 2003, 22 (1), 24–35. 10.1093/emboj/cdg014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Agresti J. J.; Antipov E.; Abate A. R.; Ahn K.; Rowat A. C.; Baret J.-C. C.; Marquez M.; Klibanov A. M.; Griffiths A. D.; Weitz D. A. Ultrahigh-Throughput Screening in Drop-Based Microfluidics for Directed Evolution. Proc. Natl. Acad. Sci. U. S. A. 2010, 107 (9), 4004–4009. 10.1073/pnas.0910781107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fallah-Araghi A.; Baret J.-C.; Ryckelynck M.; Griffiths A. D. A Completely in Vitro Ultrahigh-Throughput Droplet-Based Microfluidic Screening System for Protein Engineering and Directed Evolution. Lab Chip 2012, 12 (5), 882–891. 10.1039/c2lc21035e. [DOI] [PubMed] [Google Scholar]
- Stapleton J. A.; Swartz J. R. Development of an in Vitro Compartmentalization Screen for High-Throughput Directed Evolution of [FeFe] Hydrogenases. PLoS One 2010, 5 (12), e15275 10.1371/journal.pone.0015275. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Holstein J. M.; Gylstorff C.; Hollfelder F. Cell-Free Directed Evolution of a Protease in Microdroplets at Ultrahigh Throughput. ACS Synth. Biol. 2021, 10, 252. 10.1021/acssynbio.0c00538. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hori Y.; Kantak C.; Murray R. M.; Abate A. R. Cell-Free Extract Based Optimization of Biomolecular Circuits with Droplet Microfluidics. Lab Chip 2017, 17 (18), 3037–3042. 10.1039/C7LC00552K. [DOI] [PubMed] [Google Scholar]
- Sakamoto R.; Noireaux V.; Maeda Y. T.. Anomalous Scaling of Gene Expression in Confined Cell-Free Reactions. Sci. Rep. 2018, 8 ( (1), ). 10.1038/s41598-018-25532-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Matsuura T.; Hosoda K.; Kazuta Y.; Ichihashi N.; Suzuki H.; Yomo T. Effects of Compartment Size on the Kinetics of Intracompartmental Multimeric Protein Synthesis. ACS Synth. Biol. 2012, 1 (9), 431–437. 10.1021/sb300041z. [DOI] [PubMed] [Google Scholar]
- Kato A.; Yanagisawa M.; Sato Y. T.; Fujiwara K.; Yoshikawa K. Cell-Sized Confinement in Microspheres Accelerates the Reaction of Gene Expression. Sci. Rep. 2012, 2 (1), 283. 10.1038/srep00283. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hansen M. M. K.; Meijer L. H. H.; Spruijt E.; Maas R. J. M.; Rosquelles M. V.; Groen J.; Heus H. A.; Huck W. T. S. Macromolecular Crowding Creates Heterogeneous Environments of Gene Expression in Picolitre Droplets. Nat. Nanotechnol. 2016, 11 (2), 191–197. 10.1038/nnano.2015.243. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Olivi L.; Berger M.; Creyghton R. N. P.; De Franceschi N.; Dekker C.; Mulder B. M.; Claassens N. J.; Ten Wolde P. R.; van der Oost J. Towards a Synthetic Cell Cycle. Nat. Commun. 2021, 12 (1), 4531. 10.1038/s41467-021-24772-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Robinson A. O.; Venero O. M.; Adamala K. P. Toward Synthetic Life: Biomimetic Synthetic Cell Communication. Curr. Opin. Chem. Biol. 2021, 64, 165–173. 10.1016/j.cbpa.2021.08.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dodevski I.; Markou G. C.; Sarkar C. A. Conceptual and Methodological Advances in Cell-Free Directed Evolution. Curr. Opin. Struct. Biol. 2015, 33, 1–7. 10.1016/j.sbi.2015.04.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hanes J.; Plückthun A. In Vitro Selection and Evolution of Functional Proteins by Using Ribosome Display. Proc. Natl. Acad. Sci. U. S. A. 1997, 94 (10), 4937–4942. 10.1073/pnas.94.10.4937. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Roberts R. W.; Szostak J. W. RNA-Peptide Fusions for the in Vitro Selection of Peptides and Proteins. Proc. Natl. Acad. Sci. U. S. A. 1997, 94 (23), 12297–12302. 10.1073/pnas.94.23.12297. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Amstutz P.; Forrer P.; Zahnd C.; Plückthun A. In Vitro Display Technologies: Novel Developments and Applications. Curr. Opin. Biotechnol. 2001, 12 (4), 400–405. 10.1016/S0958-1669(00)00234-2. [DOI] [PubMed] [Google Scholar]
- Newton M.; Cabezas-Perusse Y.; Tong C. L.; Seelig B. In Vitro Selection of Peptides and Proteins - Advantages of MRNA Display. ACS Synth. Biol. 2020, 9, 181. 10.1021/acssynbio.9b00419. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cotten S. W.; Zou J.; Valencia C. A.; Liu R. Selection of Proteins with Desired Properties from Natural Proteome Libraries Using MRNA Display. Nat. Protoc. 2011, 6 (8), 1163–1182. 10.1038/nprot.2011.354. [DOI] [PubMed] [Google Scholar]
- Ju W.; Valencia C. A.; Pang H.; Ke Y.; Gao W.; Dong B.; Liu R. Proteome-Wide Identification of Family Member-Specific Natural Substrate Repertoire of Caspases. Proc. Natl. Acad. Sci. U. S. A. 2007, 104 (36), 14294–14299. 10.1073/pnas.0702251104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Du Y.; Hultquist J. F.; Zhou Q.; Olson A.; Tseng Y.; Zhang T.-H.; Hong M.; Tang K.; Chen L.; Meng X.; McGregor M. J.; Dai L.; Gong D.; Martin-Sancho L.; Chanda S.; Li X.; Bensenger S.; Krogan N. J.; Sun R. MRNA Display with Library of Even-Distribution Reveals Cellular Interactors of Influenza Virus NS1. Nat. Commun. 2020, 11 (1), 2449. 10.1038/s41467-020-16140-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gu L.; Li C.; Aach J.; Hill D. E.; Vidal M.; Church G. M. Multiplex Single-Molecule Interaction Profiling of DNA-Barcoded Proteins. Nature 2014, 515 (7528), 554–557. 10.1038/nature13761. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jermutus L.; Honegger A.; Schwesinger F.; Hanes J.; Pluckthun A. Tailoring in Vitro Evolution for Protein Affinity or Stability. Proc. Natl. Acad. Sci. U. S. A. 2001, 98 (1), 75–80. 10.1073/pnas.98.1.75. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Matsuura T.; Plückthun A. Selection Based on the Folding Properties of Proteins with Ribosome Display. FEBS Lett. 2003, 539 (1–3), 24–28. 10.1016/S0014-5793(03)00178-9. [DOI] [PubMed] [Google Scholar]
- Chaput J. C.; Szostak J. W. Evolutionary Optimization of a Nonbiological ATP Binding Protein for Improved Folding Stability. Chem. Biol. 2004, 11 (6), 865–874. 10.1016/j.chembiol.2004.04.006. [DOI] [PubMed] [Google Scholar]
- Smith M. D.; Rosenow M. A.; Wang M.; Allen J. P.; Szostak J. W.; Chaput J. C. Structural Insights into the Evolution of a Non-Biological Protein: Importance of Surface Residues in Protein Fold Optimization. PLoS One 2007, 2 (5), e467 10.1371/journal.pone.0000467. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Morelli A.; Haugner J.; Seelig B. Thermostable Artificial Enzyme Isolated by in Vitro Selection. PLoS One 2014, 9 (11), e112028. 10.1371/journal.pone.0112028. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Seelig B. MRNA Display for the Selection and Evolution of Enzymes from in Vitro-Translated Protein Libraries. Nat. Protoc. 2011, 6 (4), 540–552. 10.1038/nprot.2011.312. [DOI] [PubMed] [Google Scholar]
- Matsuura T.; Tanimura N.; Hosoda K.; Yomo T.; Shimizu Y. Reaction Dynamics Analysis of a Reconstituted Escherichia Coli Protein Translation System by Computational Modeling. Proc. Natl. Acad. Sci. U. S. A. 2017, 114 (8), E1336-E1344 10.1073/pnas.1615351114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Müller J.; Siemann-Herzberg M.; Takors R. Modeling Cell-Free Protein Synthesis Systems-Approaches and Applications. Front. Bioeng. Biotechnol. 2020, 8, 584178. 10.3389/fbioe.2020.584178. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Karzbrun E.; Shin J.; Bar-Ziv R. H.; Noireaux V. Coarse-Grained Dynamics of Protein Synthesis in a Cell-Free System. Phys. Rev. Lett. 2011, 106 (4), 048104. 10.1103/PhysRevLett.106.048104. [DOI] [PubMed] [Google Scholar]
- Stögbauer T.; Windhager L.; Zimmer R.; Rädler J. O. Experiment and Mathematical Modeling of Gene Expression Dynamics in a Cell-Free System. Integr. Biol. (Camb.) 2012, 4 (5), 494–501. 10.1039/c2ib00102k. [DOI] [PubMed] [Google Scholar]
- Horvath N.; Vilkhovoy M.; Wayman J. A.; Calhoun K.; Swartz J.; Varner J. D. Toward a Genome Scale Sequence Specific Dynamic Model of Cell-Free Protein Synthesis in Escherichia Coli. Metab. Eng. Commun. 2020, 10, e00113 10.1016/j.mec.2019.e00113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Singhal V.; Tuza Z. A.; Sun Z. Z.; Murray R. M. A MATLAB Toolbox for Modeling Genetic Circuits in Cell-Free Systems. Syn. Biol. 2021, 6, 1–14. 10.1093/synbio/ysab007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ryabova L. A.; Desplancq D.; Spirin A. S.; Plückthun A. Functional Antibody Production Using Cell-Free Translation: Effects of Protein Disulfide Isomerase and Chaperones. Nat. Biotechnol. 1997, 15 (1), 79–84. 10.1038/nbt0197-79. [DOI] [PubMed] [Google Scholar]
- Kawasaki T.; Gouda M. D.; Sawasaki T.; Takai K.; Endo Y. Efficient Synthesis of a Disulfide-Containing Protein through a Batch Cell-Free System from Wheat Germ. Eur. J. Biochem. 2003, 270 (23), 4780–4786. 10.1046/j.1432-1033.2003.03880.x. [DOI] [PubMed] [Google Scholar]
- Kim D.-M.; Swartz J. R. Efficient Production of a Bioactive, Multiple Disulfide-Bonded Protein Using Modified Extracts OfEscherichia Coli. Biotechnol. Bioeng. 2004, 85 (2), 122–129. 10.1002/bit.10865. [DOI] [PubMed] [Google Scholar]
- Yin G.; Swartz J. R. Enhancing Multiple Disulfide Bonded Protein Folding in a Cell-Free System. Biotechnol. Bioeng. 2004, 86 (2), 188–195. 10.1002/bit.10827. [DOI] [PubMed] [Google Scholar]
- Yang J.; Kanter G.; Voloshin A.; Michel-Reydellet N.; Velkeen H.; Levy R.; Swartz J. R. Rapid Expression of Vaccine Proteins for B-Cell Lymphoma in a Cell-Free System. Biotechnol. Bioeng. 2005, 89 (5), 503–511. 10.1002/bit.20283. [DOI] [PubMed] [Google Scholar]
- Ishihara G.; Goto M.; Saeki M.; Ito K.; Hori T.; Kigawa T.; Shirouzu M.; Yokoyama S. Expression of G Protein Coupled Receptors in a Cell-Free Translational System Using Detergents and Thioredoxin-Fusion Vectors. Protein Expr. Purif. 2005, 41 (1), 27–37. 10.1016/j.pep.2005.01.013. [DOI] [PubMed] [Google Scholar]
- Swartz J. Developing Cell-Free Biology for Industrial Applications. J. Ind. Microbiol. Biotechnol. 2006, 33 (7), 476–485. 10.1007/s10295-006-0127-y. [DOI] [PubMed] [Google Scholar]
- Boyer M. E.; Stapleton J. A.; Kuchenreuther J. M.; Wang C.-W.; Swartz J. R. Cell-Free Synthesis and Maturation of [FeFe] Hydrogenases. Biotechnol. Bioeng. 2008, 99 (1), 59–67. 10.1002/bit.21511. [DOI] [PubMed] [Google Scholar]
- El-Baky N. A.; Omar S. H.; Redwan E. M. The Anti-Cancer Activity of Human Consensus Interferon-Alpha Synthesized in Cell-Free System. Protein Expr. Purif. 2011, 80 (1), 61–67. 10.1016/j.pep.2011.07.003. [DOI] [PubMed] [Google Scholar]
- Matthies D.; Haberstock S.; Joos F.; Dötsch V.; Vonck J.; Bernhard F.; Meier T. Cell-Free Expression and Assembly of ATP Synthase. J. Mol. Biol. 2011, 413 (3), 593–603. 10.1016/j.jmb.2011.08.055. [DOI] [PubMed] [Google Scholar]
- Bundy B. C.; Swartz J. R. Efficient Disulfide Bond Formation in Virus-like Particles. J. Biotechnol. 2011, 154 (4), 230–239. 10.1016/j.jbiotec.2011.04.011. [DOI] [PubMed] [Google Scholar]
- Groff D.; Armstrong S.; Rivers P. J.; Zhang J.; Yang J.; Green E.; Rozzelle J.; Liang S.; Kittle J. D.; Steiner A. R.; Baliga R.; Thanos C. D.; Hallam T. J.; Sato A. K.; Yam A. Y. Engineering toward a Bacterial “Endoplasmic Reticulum” for the Rapid Expression of Immunoglobulin Proteins. MAbs 2014, 6 (3), 671–678. 10.4161/mabs.28172. [DOI] [PMC free article] [PubMed] [Google Scholar]
- He W.; Scharadin T. M.; Saldana M.; Gellner C.; Hoang-Phou S.; Takanishi C.; Hura G. L.; Tainer J. A.; Carraway K. L. 3rd; Henderson P. T.; Coleman M. A. Cell-Free Expression of Functional Receptor Tyrosine Kinases. Sci. Rep. 2015, 5 (1), 12896. 10.1038/srep12896. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sheng J.; Lei S.; Yuan L.; Feng X. Cell-Free Protein Synthesis of Norovirus Virus-like Particles. RSC Adv. 2017, 7 (46), 28837–28840. 10.1039/C7RA03742B. [DOI] [Google Scholar]
- Murakami S.; Matsumoto R.; Kanamori T. Constructive Approach for Synthesis of a Functional IgG Using a Reconstituted Cell-Free Protein Synthesis System. Sci. Rep. 2019, 9 (1), 671. 10.1038/s41598-018-36691-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hartl F. U.; Bracher A.; Hayer-Hartl M. Molecular Chaperones in Protein Folding and Proteostasis. Nature 2011, 475 (7356), 324–332. 10.1038/nature10317. [DOI] [PubMed] [Google Scholar]
- Hartl F. U. Molecular Chaperones in Cellular Protein Folding. Nature 1996, 381 (6583), 571–579. 10.1038/381571a0. [DOI] [PubMed] [Google Scholar]
- Waudby C. A.; Dobson C. M.; Christodoulou J. Nature and Regulation of Protein Folding on the Ribosome. Trends Biochem. Sci. 2019, 44, 914–926. 10.1016/j.tibs.2019.06.008. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yan X.; Hoek T. A.; Vale R. D.; Tanenbaum M. E. Dynamics of Translation of Single MRNA Molecules in Vivo. Cell 2016, 165 (4), 976–989. 10.1016/j.cell.2016.04.034. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Welsh J. P.; Bonomo J.; Swartz J. R. Localization of BiP to Translating Ribosomes Increases Soluble Accumulation of Secreted Eukaryotic Proteins in an Escherichia Coli Cell-Free System. Biotechnol. Bioeng. 2011, 108 (8), 1739–1748. 10.1002/bit.23111. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sasaki Y.; Asayama W.; Niwa T.; Sawada S.-I.; Ueda T.; Taguchi H.; Akiyoshi K. Amphiphilic Polysaccharide Nanogels as Artificial Chaperones in Cell-Free Protein Synthesis. Macromol. Biosci. 2011, 11 (6), 814–820. 10.1002/mabi.201000457. [DOI] [PubMed] [Google Scholar]
- Lim H. J.; Lee K.-H.; Kim D.-M. Rapid Determination of Effective Folding Agents by Sequential Cell-Free Protein Synthesis. Biochem. Eng. J. 2018, 138, 106–110. 10.1016/j.bej.2018.07.014. [DOI] [Google Scholar]
- Niwa T.; Kanamori T.; Ueda T.; Taguchi H. Global Analysis of Chaperone Effects Using a Reconstituted Cell-Free Translation System. Proc. Natl. Acad. Sci. U. S. A. 2012, 109 (23), 8937–8942. 10.1073/pnas.1201380109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mol. Biol. Cell, 4th ed.; Alberts B., et al., Eds.; CRC Press: Boca Raton, FL, 2002. [Google Scholar]
- Sevier C. S.; Kaiser C. A. Formation and Transfer of Disulphide Bonds in Living Cells. Nat. Rev. Mol. Cell Biol. 2002, 3 (11), 836–847. 10.1038/nrm954. [DOI] [PubMed] [Google Scholar]
- Stech M.; Kubick S. Cell-Free Synthesis Meets Antibody Production: A Review. Antibodies 2015, 4 (1), 12–33. 10.3390/antib4010012. [DOI] [Google Scholar]
- Sachse R.; Dondapati S. K.; Fenz S. F.; Schmidt T.; Kubick S. Membrane Protein Synthesis in Cell-Free Systems: From Bio-Mimetic Systems to Bio-Membranes. FEBS Lett. 2014, 588 (17), 2774–2781. 10.1016/j.febslet.2014.06.007. [DOI] [PubMed] [Google Scholar]
- Kruyer N. S.; Sugianto W.; Tickman B. I.; Alba Burbano D.; Noireaux V.; Carothers J. M.; Peralta-Yahya P. Membrane Augmented Cell-Free Systems: A New Frontier in Biotechnology. ACS Synth. Biol. 2021, 10 (4), 670–681. 10.1021/acssynbio.0c00625. [DOI] [PubMed] [Google Scholar]
- Müller M.; Blobel G. In Vitro Translocation of Bacterial Proteins across the Plasma Membrane of Escherichia Coli. Proc. Natl. Acad. Sci. U. S. A. 1984, 81 (23), 7421–7425. 10.1073/pnas.81.23.7421. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wuu J. J.; Swartz J. R. High Yield Cell-Free Production of Integral Membrane Proteins without Refolding or Detergents. Biochim. Biophys. Acta 2008, 1778 (5), 1237–1250. 10.1016/j.bbamem.2008.01.023. [DOI] [PubMed] [Google Scholar]
- Sachse R.; Wüstenhagen D.; Šamalíková M.; Gerrits M.; Bier F. F.; Kubick S. Synthesis of Membrane Proteins in Eukaryotic Cell-free Systems. Eng. Life Sci. 2013, 13 (1), 39–48. 10.1002/elsc.201100235. [DOI] [Google Scholar]
- Kuruma Y.; Nishiyama K.-I.; Shimizu Y.; Müller M.; Ueda T. Development of a Minimal Cell-Free Translation System for the Synthesis of Presecretory and Integral Membrane Proteins. Biotechnol. Prog. 2005, 21 (4), 1243–1251. 10.1021/bp049553u. [DOI] [PubMed] [Google Scholar]
- Joseph S. K.; Boehning D.; Pierson S.; Nicchitta C. V. Membrane Insertion, Glycosylation, and Oligomerization of Inositol Trisphosphate Receptors in a Cell-Free Translation System*. J. Biol. Chem. 1997, 272 (3), 1579–1588. 10.1074/jbc.272.3.1579. [DOI] [PubMed] [Google Scholar]
- Moghimianavval H.; Hsu Y.-Y.; Groaz A.; Liu A. P.. In Vitro Reconstitution Platforms of Mammalian Cell-Free Expressed Membrane ProteinsMembrane Proteins. In Cell-Free Gene Expression: Methods and Protocols; Karim A. S., Jewett M. C., Eds.; Springer US: New York, NY, 2022; pp 105–120. 10.1007/978-1-0716-1998-8_6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Katzen F.; Kudlicki W. Efficient Generation of Insect-Based Cell-Free Translation Extracts Active in Glycosylation and Signal Sequence Processing. J. Biotechnol. 2006, 125 (2), 194–197. 10.1016/j.jbiotec.2006.03.002. [DOI] [PubMed] [Google Scholar]
- Dondapati S. K.; Kreir M.; Quast R. B.; Wüstenhagen D. A.; Brüggemann A.; Fertig N.; Kubick S. Membrane Assembly of the Functional KcsA Potassium Channel in a Vesicle-Based Eukaryotic Cell-Free Translation System. Biosens. Bioelectron. 2014, 59, 174–183. 10.1016/j.bios.2014.03.004. [DOI] [PubMed] [Google Scholar]
- Hovijitra N. T.; Wuu J. J.; Peaker B.; Swartz J. R. Cell-Free Synthesis of Functional Aquaporin Z in Synthetic Liposomes. Biotechnol. Bioeng. 2009, 104 (1), 40–49. 10.1002/bit.22385. [DOI] [PubMed] [Google Scholar]
- Jacobs M. L.; Boyd M. A.; Kamat N. P. Diblock Copolymers Enhance Folding of a Mechanosensitive Membrane Protein during Cell-Free Expression. Proc. Natl. Acad. Sci. U. S. A. 2019, 116 (10), 4031–4036. 10.1073/pnas.1814775116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ridder A. N. J. A.; van de Hoef W.; Stam J.; Kuhn A.; de Kruijff B.; Killian J. A. Importance of Hydrophobic Matching for Spontaneous Insertion of a Single-Spanning Membrane Protein. Biochemistry 2002, 41 (15), 4946–4952. 10.1021/bi0158674. [DOI] [PubMed] [Google Scholar]
- Kalmbach R.; Chizhov I.; Schumacher M. C.; Friedrich T.; Bamberg E.; Engelhard M. Functional Cell-Free Synthesis of a Seven Helix Membrane Protein: In Situ Insertion of Bacteriorhodopsin into Liposomes. J. Mol. Biol. 2007, 371 (3), 639–648. 10.1016/j.jmb.2007.05.087. [DOI] [PubMed] [Google Scholar]
- Harris N. J.; Pellowe G. A.; Booth P. J. Cell-Free Expression Tools to Study Co-Translational Folding of Alpha Helical Membrane Transporters. Sci. Rep. 2020, 10 (1), 9125. 10.1038/s41598-020-66097-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Blackholly L. R.; Harris N. J.; Findlay H. E.; Booth P. J. Cell-Free Expression to Probe Co-Translational Insertion of an Alpha Helical Membrane Protein. Front Mol. Biosci 2022, 9, 795212. 10.3389/fmolb.2022.795212. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schoenmakers L. L. J.; Yewdall N. A.; Lu T.; André A. A. M.; Nelissen F. H. T.; Spruijt E.; Huck W. T. S. In Vitro Transcription-Translation in an Artificial Biomolecular Condensate. ACS Synth. Biol. 2023, 12 (7), 2004–2014. 10.1021/acssynbio.3c00069. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bayburt T. H.; Sligar S. G. Membrane Protein Assembly into Nanodiscs. FEBS Lett. 2010, 584 (9), 1721–1727. 10.1016/j.febslet.2009.10.024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jaroentomeechai T.; Zheng X.; Hershewe J.; Stark J. C.; Jewett M. C.; DeLisa M. P.. Chapter Three - A Pipeline for Studying and Engineering Single-Subunit Oligosaccharyltransferases. In Methods in Enzymology; Imperiali B., Ed.; Academic Press, 2017; Vol. 597, pp 55–81. 10.1016/bs.mie.2017.07.011. [DOI] [PubMed] [Google Scholar]
- Henrich E.; Löhr F.; Pawlik G.; Peetz O.; Dötsch V.; Morgner N.; de Kroon A. I.; Bernhard F. Lipid Conversion by Cell-Free Synthesized Phospholipid Methyltransferase Opi3 in Defined Nanodisc Membranes Supports an in Trans Mechanism. Biochemistry 2018, 57 (40), 5780–5784. 10.1021/acs.biochem.8b00807. [DOI] [PubMed] [Google Scholar]
- Gessesse B.; Nagaike T.; Nagata K.; Shimizu Y.; Ueda T. G-Protein Coupled Receptor Protein Synthesis on a Lipid Bilayer Using a Reconstituted Cell-Free Protein Synthesis System. Life 2018, 8 (4), 54. 10.3390/life8040054. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Henrich E.; Dötsch V.; Bernhard F.. Chapter Sixteen - Screening for Lipid Requirements of Membrane Proteins by Combining Cell-Free Expression with Nanodiscs. In Methods in Enzymology; Shukla A. K., Ed.; Academic Press, 2015; Vol. 556, pp 351–369. 10.1016/bs.mie.2014.12.016. [DOI] [PubMed] [Google Scholar]
- Berrier C.; Park K.-H.; Abes S.; Bibonne A.; Betton J.-M.; Ghazi A. Cell-Free Synthesis of a Functional Ion Channel in the Absence of a Membrane and in the Presence of Detergent. Biochemistry 2004, 43 (39), 12585–12591. 10.1021/bi049049y. [DOI] [PubMed] [Google Scholar]
- Lyukmanova E. N.; Shenkarev Z. O.; Khabibullina N. F.; Kopeina G. S.; Shulepko M. A.; Paramonov A. S.; Mineev K. S.; Tikhonov R. V.; Shingarova L. N.; Petrovskaya L. E.; Dolgikh D. A.; Arseniev A. S.; Kirpichnikov M. P. Lipid-Protein Nanodiscs for Cell-Free Production of Integral Membrane Proteins in a Soluble and Folded State: Comparison with Detergent Micelles, Bicelles and Liposomes. Biochim. Biophys. Acta 2012, 1818 (3), 349–358. 10.1016/j.bbamem.2011.10.020. [DOI] [PubMed] [Google Scholar]
- Mezhyrova J.; Mörs K.; Glaubitz C.; Dötsch V.; Bernhard F.. Applications of Cell-Free Synthesized Membrane Protein Precipitates. In Insoluble Proteins: Methods and Protocols; Garcia Fruitós E., Arís Giralt A., Eds.; Springer US: New York, NY, 2022; pp 245–266. 10.1007/978-1-0716-1859-2_15. [DOI] [PubMed] [Google Scholar]
- Boland C.; Li D.; Shah S. T. A.; Haberstock S.; Dötsch V.; Bernhard F.; Caffrey M. Cell-Free Expression and in Meso Crystallisation of an Integral Membrane Kinase for Structure Determination. Cell. Mol. Life Sci. 2014, 71 (24), 4895–4910. 10.1007/s00018-014-1655-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Shimono K.; Goto M.; Kikukawa T.; Miyauchi S.; Shirouzu M.; Kamo N.; Yokoyama S. Production of Functional Bacteriorhodopsin by an Escherichia Coli Cell-Free Protein Synthesis System Supplemented with Steroid Detergent and Lipid. Protein Sci. 2009, 18 (10), 2160–2171. 10.1002/pro.230. [DOI] [PMC free article] [PubMed] [Google Scholar]
- One-pot system for synthesis, assembly, and display of functional single-span membrane proteins on oil-water interfaces. PNAS. https://www.pnas.org/doi/full/10.1073/pnas.1504992113 (accessed 2022-04-13). [DOI] [PMC free article] [PubMed]
- Exterkate M.; Caforio A.; Stuart M. C. A.; Driessen A. J. M. Growing Membranes In Vitro by Continuous Phospholipid Biosynthesis from Free Fatty Acids. ACS Synth. Biol. 2018, 7 (1), 153–165. 10.1021/acssynbio.7b00265. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Scott A.; Noga M. J.; de Graaf P.; Westerlaken I.; Yildirim E.; Danelon C. Cell-Free Phospholipid Biosynthesis by Gene-Encoded Enzymes Reconstituted in Liposomes. PLoS One 2016, 11 (10), e0163058 10.1371/journal.pone.0163058. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Abu-Qarn M.; Eichler J.; Sharon N. Not Just for Eukarya Anymore: Protein Glycosylation in Bacteria and Archaea. Curr. Opin. Struct. Biol. 2008, 18 (5), 544–550. 10.1016/j.sbi.2008.06.010. [DOI] [PubMed] [Google Scholar]
- Kay E.; Cuccui J.; Wren B. W. Recent Advances in the Production of Recombinant Glycoconjugate Vaccines. NPJ. Vaccines 2019, 4, 16. 10.1038/s41541-019-0110-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schwarz F.; Aebi M. Mechanisms and Principles of N-Linked Protein Glycosylation. Curr. Opin. Struct. Biol. 2011, 21 (5), 576–582. 10.1016/j.sbi.2011.08.005. [DOI] [PubMed] [Google Scholar]
- Jefferis R. Glycosylation as a Strategy to Improve Antibody-Based Therapeutics. Nat. Rev. Drug Discovery 2009, 8 (3), 226–234. 10.1038/nrd2804. [DOI] [PubMed] [Google Scholar]
- Rothblatt J. A.; Meyer D. I. Secretion in Yeast: Reconstitution of the Translocation and Glycosylation of Alpha-Factor and Invertase in a Homologous Cell-Free System. Cell 1986, 44 (4), 619–628. 10.1016/0092-8674(86)90271-0. [DOI] [PubMed] [Google Scholar]
- Bailey C. A.; Gerber L.; Howard A. D.; Udenfriend S. Processing at the Carboxyl Terminus of Nascent Placental Alkaline Phosphatase in a Cell-Free System: Evidence for Specific Cleavage of a Signal Peptide. Proc. Natl. Acad. Sci. U. S. A. 1989, 86 (1), 22–26. 10.1073/pnas.86.1.22. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thoring L.; Wüstenhagen D. A.; Borowiak M.; Stech M.; Sonnabend A.; Kubick S. Cell-Free Systems Based on CHO Cell Lysates: Optimization Strategies, Synthesis of “Difficult-to-Express” Proteins and Future Perspectives. PLoS One 2016, 11 (9), e0163670 10.1371/journal.pone.0163670. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Quast R. B.; Sonnabend A.; Stech M.; Wüstenhagen D. A.; Kubick S. High-Yield Cell-Free Synthesis of Human EGFR by IRES-Mediated Protein Translation in a Continuous Exchange Cell-Free Reaction Format. Sci. Rep. 2016, 6, 30399. 10.1038/srep30399. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gurramkonda C.; Rao A.; Borhani S.; Pilli M.; Deldari S.; Ge X.; Pezeshk N.; Han T.-C.; Tolosa M.; Kostov Y.; Tolosa L.; Wood D. W.; Vattem K.; Frey D. D.; Rao G. Improving the Recombinant Human Erythropoietin Glycosylation Using Microsome Supplementation in CHO Cell-Free System. Biotechnol. Bioeng. 2018, 115 (5), 1253–1264. 10.1002/bit.26554. [DOI] [PubMed] [Google Scholar]
- Kightlinger W.; Warfel K. F.; DeLisa M. P.; Jewett M. C. Synthetic Glycobiology: Parts, Systems, and Applications. ACS Synth. Biol. 2020, 9 (7), 1534–1562. 10.1021/acssynbio.0c00210. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wacker M.; Linton D.; Hitchen P. G.; Nita-Lazar M.; Haslam S. M.; North S. J.; Panico M.; Morris H. R.; Dell A.; Wren B. W.; Aebi M. N-Linked Glycosylation in Campylobacter Jejuni and Its Functional Transfer into E. Coli. Science 2002, 298 (5599), 1790–1793. 10.1126/science.298.5599.1790. [DOI] [PubMed] [Google Scholar]
- Feldman M. F.; Wacker M.; Hernandez M.; Hitchen P. G.; Marolda C. L.; Kowarik M.; Morris H. R.; Dell A.; Valvano M. A.; Aebi M. Engineering N-Linked Protein Glycosylation with Diverse O Antigen Lipopolysaccharide Structures in Escherichia Coli. Proc. Natl. Acad. Sci. U. S. A. 2005, 102 (8), 3016–3021. 10.1073/pnas.0500044102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guarino C.; DeLisa M. P. A Prokaryote-Based Cell-Free Translation System That Efficiently Synthesizes Glycoproteins. Glycobiology 2012, 22 (5), 596–601. 10.1093/glycob/cwr151. [DOI] [PubMed] [Google Scholar]
- Baker J. L.; Çelik E.; DeLisa M. P. Expanding the Glycoengineering Toolbox: The Rise of Bacterial N-Linked Protein Glycosylation. Trends Biotechnol. 2013, 31 (5), 313–323. 10.1016/j.tibtech.2013.03.003. [DOI] [PubMed] [Google Scholar]
- Aquino A. K.; Manzer Z. A.; Daniel S.; DeLisa M. P. Glycosylation-on-a-Chip: A Flow-Based Microfluidic System for Cell-Free Glycoprotein Biosynthesis. Front Mol. Biosci 2021, 8, 782905. 10.3389/fmolb.2021.782905. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Natarajan A.; Jaroentomeechai T.; Cabrera-Sánchez M.; Mohammed J. C.; Cox E. C.; Young O.; Shajahan A.; Vilkhovoy M.; Vadhin S.; Varner J. D.; Azadi P.; DeLisa M. P. Engineering Orthogonal Human O-Linked Glycoprotein Biosynthesis in Bacteria. Nat. Chem. Biol. 2020, 16 (10), 1062–1070. 10.1038/s41589-020-0595-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kightlinger W.; Lin L.; Rosztoczy M.; Li W.; DeLisa M. P.; Mrksich M.; Jewett M. C. Design of Glycosylation Sites by Rapid Synthesis and Analysis of Glycosyltransferases. Nat. Chem. Biol. 2018, 14 (6), 627–635. 10.1038/s41589-018-0051-2. [DOI] [PubMed] [Google Scholar]
- Techner J.-M.; Kightlinger W.; Lin L.; Hershewe J.; Ramesh A.; DeLisa M. P.; Jewett M. C.; Mrksich M. High-Throughput Synthesis and Analysis of Intact Glycoproteins Using SAMDI-MS. Anal. Chem. 2020, 92 (2), 1963–1971. 10.1021/acs.analchem.9b04334. [DOI] [PubMed] [Google Scholar]
- Lin L.; Kightlinger W.; Prabhu S. K.; Hockenberry A. J.; Li C.; Wang L.-X.; Jewett M. C.; Mrksich M. Sequential Glycosylation of Proteins with Substrate-Specific N-Glycosyltransferases. ACS Cent Sci. 2020, 6 (2), 144–154. 10.1021/acscentsci.9b00021. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kinoshita-Kikuta E.; Kinoshita E.; Suga M.; Higashida M.; Yamane Y.; Nakamura T.; Koike T. Characterization of Phosphorylation Status and Kinase Activity of Src Family Kinases Expressed in Cell-Based and Cell-Free Protein Expression Systems. Biomolecules 2021, 11 (10), 1448. 10.3390/biom11101448. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Safer B.; Jagus R. Control of EIF-2 Phosphatase Activity in Rabbit Reticulocyte Lysate. Proc. Natl. Acad. Sci. U. S. A. 1979, 76 (3), 1094–1098. 10.1073/pnas.76.3.1094. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Suzuki T.; Ito M.; Ezure T.; Shikata M.; Ando E.; Utsumi T.; Tsunasawa S.; Nishimura O. Protein Prenylation in an Insect Cell-Free Protein Synthesis System and Identification of Products by Mass Spectrometry. Proteomics 2007, 7 (12), 1942–1950. 10.1002/pmic.200700237. [DOI] [PubMed] [Google Scholar]
- Hancock J. F.[7] Reticulocyte Lysate Assay for in Vitro Translation and Posttranslational Modification of Ras Proteins. In Methods in Enzymology; Academic Press, 1995; Vol. 255, pp 60–65. 10.1016/S0076-6879(95)55009-7. [DOI] [PubMed] [Google Scholar]
- Suzuki T.; Ito M.; Ezure T.; Shikata M.; Ando E.; Utsumi T.; Tsunasawa S.; Nishimura O. N-Terminal Protein Modifications in an Insect Cell-Free Protein Synthesis System and Their Identification by Mass Spectrometry. Proteomics 2006, 6 (16), 4486–4495. 10.1002/pmic.200600126. [DOI] [PubMed] [Google Scholar]
- Gibbs P. E.; Zouzias D. C.; Freedberg I. M. Differential Post-Translational Modification of Human Type I Keratins Synthesized in a Rabbit Reticulocyte Cell-Free System. Biochim. Biophys. Acta 1985, 824 (3), 247–255. 10.1016/0167-4781(85)90055-7. [DOI] [PubMed] [Google Scholar]
- Suzuki T.; Ezure T.; Ando E.; Nishimura O.; Utsumi T.; Tsunasawa S. Preparation of Ubiquitin-Conjugated Proteins Using an Insect Cell-Free Protein Synthesis System. J. Biotechnol. 2010, 145 (1), 73–78. 10.1016/j.jbiotec.2009.10.009. [DOI] [PubMed] [Google Scholar]
- Kai L.; Sonal; Heermann T.; Schwille P. One-Pot Synthesis of Prenylated Proteins Utilizing E. Coli Cell-Free Expression. bioRxiv 2022, 10.1101/2022.03.03.482798. [DOI] [Google Scholar]
- Oza J. P.; Aerni H. R.; Pirman N. L.; Barber K. W.; Ter Haar C. M.; Rogulina S.; Amrofell M. B.; Isaacs F. J.; Rinehart J.; Jewett M. C.. Robust Production of Recombinant Phosphoproteins Using Cell-Free Protein Synthesis. Nat. Commun. 2015, 6. 10.1038/ncomms9168. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jin X.; Hong S. H. Cell-Free Protein Synthesis for Producing ‘Difficult-to-Express’ Proteins. Biochem. Eng. J. 2018, 138, 156–164. 10.1016/j.bej.2018.07.013. [DOI] [Google Scholar]
- Lüddecke T.; Paas A.; Talmann L.; Kirchhoff K. N.; von Reumont B. M.; Billion A.; Timm T.; Lochnit G.; Vilcinskas A. A Spider Toxin Exemplifies the Promises and Pitfalls of Cell-Free Protein Production for Venom Biodiscovery. Toxins (Basel) 2021, 13 (8), 575. 10.3390/toxins13080575. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ramm F.; Stech M.; Zemella A.; Frentzel H.; Kubick S. The Pore-Forming Hemolysin BL Enterotoxin from Bacillus Cereus: Subunit Interactions in Cell-Free Systems. Toxins (Basel) 2021, 13 (11), 807. 10.3390/toxins13110807. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Eichenlaub R. Synthesis of Colicin E1 in a Cell-Free System. J. Bacteriol. 1974, 120 (3), 1476–1477. 10.1128/jb.120.3.1476-1477.1974. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chalmeau J.; Monina N.; Shin J.; Vieu C.; Noireaux V. α-Hemolysin Pore Formation into a Supported Phospholipid Bilayer Using Cell-Free Expression. Biochim. Biophys. Acta 2011, 1808 (1), 271–278. 10.1016/j.bbamem.2010.07.027. [DOI] [PubMed] [Google Scholar]
- Jin X.; Kightlinger W.; Kwon Y.-C.; Hong S. H. Rapid Production and Characterization of Antimicrobial Colicins Using Escherichia Coli-Based Cell-Free Protein Synthesis. Synth. Biol. 2018, 3 (1), 1–11. 10.1093/synbio/ysy004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ramm F.; Jack L.; Kaser D.; Schloßhauer J. L.; Zemella A.; Kubick S. Cell-Free Systems Enable the Production of AB5 Toxins for Diagnostic Applications. Toxins (Basel) 2022, 14 (4), 233. 10.3390/toxins14040233. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cintas L. M.; Casaus P.; Holo H.; Hernandez P. E.; Nes I. F.; Håvarstein L. S. Enterocins L50A and L50B, Two Novel Bacteriocins from Enterococcus Faecium L50, Are Related to Staphylococcal Hemolysins. J. Bacteriol. 1998, 180 (8), 1988–1994. 10.1128/JB.180.8.1988-1994.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Orth J. H. C.; Schorch B.; Boundy S.; Ffrench-Constant R.; Kubick S.; Aktories K. Cell-Free Synthesis and Characterization of a Novel Cytotoxic Pierisin-like Protein from the Cabbage Butterfly Pieris Rapae. Toxicon 2011, 57 (2), 199–207. 10.1016/j.toxicon.2010.11.011. [DOI] [PubMed] [Google Scholar]
- Krinsky N.; Kaduri M.; Shainsky-Roitman J.; Goldfeder M.; Ivanir E.; Benhar I.; Shoham Y.; Schroeder A. A Simple and Rapid Method for Preparing a Cell-Free Bacterial Lysate for Protein Synthesis. PLoS One 2016, 11 (10), e0165137. 10.1371/journal.pone.0165137. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smith M. T.; Varner C. T.; Bush D. B.; Bundy B. C. The Incorporation of the A2 Protein to Produce Novel Qβ Virus-like Particles Using Cell-Free Protein Synthesis. Biotechnol. Prog. 2012, 28 (2), 549–555. 10.1002/btpr.744. [DOI] [PubMed] [Google Scholar]
- Bechlars S.; Wüstenhagen D. A.; Drägert K.; Dieckmann R.; Strauch E.; Kubick S. Cell-Free Synthesis of Functional Thermostable Direct Hemolysins of Vibrio Parahaemolyticus. Toxicon 2013, 76, 132–142. 10.1016/j.toxicon.2013.09.012. [DOI] [PubMed] [Google Scholar]
- Dondapati S. K.; Wüstenhagen D. A.; Strauch E.; Kubick S. Cell-Free Production of Pore Forming Toxins: Functional Analysis of Thermostable Direct Hemolysin from Vibrio Parahaemolyticus. Eng. Life Sci. 2018, 18 (2), 140–148. 10.1002/elsc.201600259. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Quast R. B.; Mrusek D.; Hoffmeister C.; Sonnabend A.; Kubick S. Cotranslational Incorporation of Non-Standard Amino Acids Using Cell-Free Protein Synthesis. FEBS Lett. 2015, 589 (15), 1703–1712. 10.1016/j.febslet.2015.04.041. [DOI] [PubMed] [Google Scholar]
- Dumas A.; Lercher L.; Spicer C. D.; Davis B. G. Designing Logical Codon Reassignment-Expanding the Chemistry in Biology. Chem. Sci. 2015, 6 (1), 50–69. 10.1039/C4SC01534G. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Young D. D.; Schultz P. G. Playing with the Molecules of Life. ACS Chem. Biol. 2018, 13, 854–870. 10.1021/acschembio.7b00974. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Davis L.; Chin J. W. Designer Proteins: Applications of Genetic Code Expansion in Cell Biology. Nat. Rev. Mol. Cell Biol. 2012, 13 (3), 168–182. 10.1038/nrm3286. [DOI] [PubMed] [Google Scholar]
- Birch-Price Z.; Taylor C. J.; Ortmayer M.; Green A. P.. Engineering Enzyme Activity Using an Expanded Amino Acid Alphabet. Protein Eng. Des. Sel. 2023, 36. 10.1093/protein/gzac013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Johnson J. A.; Lu Y. Y.; Van Deventer J. A.; Tirrell D. A. Residue-Specific Incorporation of Non-Canonical Amino Acids into Proteins: Recent Developments and Applications. Curr. Opin. Chem. Biol. 2010, 14 (6), 774–780. 10.1016/j.cbpa.2010.09.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chapeville F.; Lipmann F.; von Ehrenstein G.; Weisblum B.; Ray W. J.; Benzer S. On the Role of Soluble Ribonucleic Acid in Coding for Amino Acids. Proc. Natl. Acad. Sci. U. S. A. 1962, 48 (1), 1086–1092. 10.1073/pnas.48.6.1086. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fahnestock S.; Rich A. Ribosome-Catalyzed Polyester Formation. Science 1971, 173 (3994), 340–343. 10.1126/science.173.3994.340. [DOI] [PubMed] [Google Scholar]
- Merryman C.; Green R. Transformation of Aminoacyl TRNAs for the in Vitro Selection of “Drug-like” Molecules. Chem. Biol. 2004, 11 (4), 575–582. 10.1016/j.chembiol.2004.03.009. [DOI] [PubMed] [Google Scholar]
- Heckler T. G.; Chang L. H.; Zama Y.; Naka T.; Chorghade M. S.; Hecht S. M. T4 RNA Ligase Mediated Preparation of Novel “Chemically Misacylated” TRNAPheS. Biochemistry 1984, 23 (7), 1468–1473. 10.1021/bi00302a020. [DOI] [PubMed] [Google Scholar]
- Robertson S. A.; Ellman J. A.; Schultz P. G. A General and Efficient Route for Chemical Aminoacylation of Transfer RNAs. J. Am. Chem. Soc. 1991, 113 (7), 2722–2729. 10.1021/ja00007a055. [DOI] [Google Scholar]
- Kwiatkowski M.; Wang J.; Forster A. C. Facile Synthesis of N-Acyl-Aminoacyl-PCpA for Preparation of Mischarged Fully Ribo TRNA. Bioconjugate Chem. 2014, 25 (11), 2086–2091. 10.1021/bc500441b. [DOI] [PubMed] [Google Scholar]
- Vargas-Rodriguez O.; Sevostyanova A.; Söll D.; Crnković A. Upgrading Aminoacyl-TRNA Synthetases for Genetic Code Expansion. Curr. Opin. Chem. Biol. 2018, 46, 115–122. 10.1016/j.cbpa.2018.07.014. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Murakami H.; Ohta A.; Ashigai H.; Suga H. A Highly Flexible TRNA Acylation Method for Non-Natural Polypeptide Synthesis. Nat. Methods 2006, 3 (5), 357–359. 10.1038/nmeth877. [DOI] [PubMed] [Google Scholar]
- Goto Y.; Katoh T.; Suga H. Flexizymes for Genetic Code Reprogramming. Nat. Protoc. 2011, 6 (6), 779–790. 10.1038/nprot.2011.331. [DOI] [PubMed] [Google Scholar]
- Katoh T.; Suga H. Ribosomal Elongation of Aminobenzoic Acid Derivatives. J. Am. Chem. Soc. 2020, 142 (39), 16518–16522. 10.1021/jacs.0c05765. [DOI] [PubMed] [Google Scholar]
- Katoh T.; Suga H. Ribosomal Elongation of Cyclic γ-Amino Acids Using a Reprogrammed Genetic Code. J. Am. Chem. Soc. 2020, 142, 4965–4969. 10.1021/jacs.9b12280. [DOI] [PubMed] [Google Scholar]
- Katoh T.; Suga H. Ribosomal Incorporation of Consecutive β-Amino Acids. J. Am. Chem. Soc. 2018, 140 (38), 12159–12167. 10.1021/jacs.8b07247. [DOI] [PubMed] [Google Scholar]
- Katoh T.; Sengoku T.; Hirata K.; Ogata K.; Suga H. Ribosomal Synthesis and de Novo Discovery of Bioactive Foldamer Peptides Containing Cyclic β-Amino Acids. Nat. Chem. 2020, 12, 1081. 10.1038/s41557-020-0525-1. [DOI] [PubMed] [Google Scholar]
- Katoh T.; Suga H. Consecutive Ribosomal Incorporation of α-Aminoxy/α-Hydrazino Acids with l/d-Configurations into Nascent Peptide Chains. J. Am. Chem. Soc. 2021, 143 (45), 18844–18848. 10.1021/jacs.1c09270. [DOI] [PubMed] [Google Scholar]
- Maini R.; Kimura H.; Takatsuji R.; Katoh T.; Goto Y.; Suga H. Ribosomal Formation of Thioamide Bonds in Polypeptide Synthesis. J. Am. Chem. Soc. 2019, 141 (51), 20004–20008. 10.1021/jacs.9b11097. [DOI] [PubMed] [Google Scholar]
- Ieong K.-W.; Pavlov M. Y.; Kwiatkowski M.; Ehrenberg M.; Forster A. C. A TRNA Body with High Affinity for EF-Tu Hastens Ribosomal Incorporation of Unnatural Amino Acids. RNA 2014, 20 (5), 632–643. 10.1261/rna.042234.113. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yamaguchi A.; Iraha F.; Ohtake K.; Sakamoto K. Pyrrolysyl-TRNA Synthetase with a Unique Architecture Enhances the Availability of Lysine Derivatives in Synthetic Genetic Codes. Molecules 2018, 23 (10), 2460. 10.3390/molecules23102460. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Uyeda A.; Watanabe T.; Kato Y.; Watanabe H.; Yomo T.; Hohsaka T.; Matsuura T. Liposome-Based in Vitro Evolution of Aminoacyl-TRNA Synthetase for Enhanced Pyrrolysine Derivative Incorporation. Chembiochem 2015, 16 (12), 1797–1802. 10.1002/cbic.201500174. [DOI] [PubMed] [Google Scholar]
- Josephson K.; Hartman M. C. T.; Szostak J. W. Ribosomal Synthesis of Unnatural Peptides. J. Am. Chem. Soc. 2005, 127 (33), 11727–11735. 10.1021/ja0515809. [DOI] [PubMed] [Google Scholar]
- Iqbal E. S.; Dods K. K.; Hartman M. C. T. Ribosomal Incorporation of Backbone Modified Amino Acids: Via an Editing-Deficient Aminoacyl-TRNA Synthetase. Org. Biomol. Chem. 2018, 16 (7), 1073–1078. 10.1039/C7OB02931D. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goto Y.; Ohta A.; Sako Y.; Yamagishi Y.; Murakami H.; Suga H. Reprogramming the Translation Initiation for the Synthesis of Physiologically Stable Cyclic Peptides. ACS Chem. Biol. 2008, 3 (2), 120–129. 10.1021/cb700233t. [DOI] [PubMed] [Google Scholar]
- Odom O. W.; Kudlicki W.; Hardesty B. In Vitro Engineering Using Acyl-Derivatized TRNAs. Methods Mol. Biol. 1998, 77, 93–103. 10.1385/0-89603-397-X:93. [DOI] [PubMed] [Google Scholar]
- Tsiamantas C.; Rogers J. M.; Suga H. Initiating Ribosomal Peptide Synthesis with Exotic Building Blocks. Chem. Commun. 2020, 56 (31), 4265–4272. 10.1039/D0CC01291B. [DOI] [PubMed] [Google Scholar]
- Cload S. T.; Liu D. R.; Froland W. A.; Schultz P. G. Development of Improved TRNAs for in Vitro Biosynthesis of Proteins Containing Unnatural Amino Acids. Chem. Biol. 1996, 3 (12), 1033–1038. 10.1016/S1074-5521(96)90169-6. [DOI] [PubMed] [Google Scholar]
- Murakami H.; Kourouklis D.; Suga H. Using a Solid-Phase Ribozyme Aminoacylation System to Reprogram the Genetic Code. Chem. Biol. 2003, 10 (11), 1077–1084. 10.1016/j.chembiol.2003.10.010. [DOI] [PubMed] [Google Scholar]
- Iwane Y.; Kimura H.; Katoh T.; Suga H. Uniform Affinity-Tuning of N-Methyl-Aminoacyl-TRNAs to EF-Tu Enhances Their Multiple Incorporation. Nucleic Acids Res. 2021, 49 (19), 10807–10817. 10.1093/nar/gkab288. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Katoh T.; Tajima K.; Suga H. Consecutive Elongation of D-Amino Acids in Translation. Cell Chemical Biology 2017, 24 (1), 46–54. 10.1016/j.chembiol.2016.11.012. [DOI] [PubMed] [Google Scholar]
- Katoh T.; Wohlgemuth I.; Nagano M.; Rodnina M. V.; Suga H. Essential Structural Elements in TRNA(Pro) for EF-P-Mediated Alleviation of Translation Stalling. Nat. Commun. 2016, 7, 11657. 10.1038/ncomms11657. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Katoh T.; Iwane Y.; Suga H. Logical Engineering of D-Arm and T-Stem of TRNA That Enhances D-Amino Acid Incorporation. Nucleic Acids Res. 2017, 45 (22), 12601–12610. 10.1093/nar/gkx1129. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chemla Y.; Ozer E.; Schlesinger O.; Noireaux V.; Alfonta L. Genetically Expanded Cell-Free Protein Synthesis Using Endogenous Pyrrolysyl Orthogonal Translation System. Biotechnol. Bioeng. 2015, 112 (8), 1663–1672. 10.1002/bit.25587. [DOI] [PubMed] [Google Scholar]
- Ozer E.; Chemla Y.; Schlesinger O.; Aviram H. Y.; Riven I.; Haran G.; Alfonta L. In Vitro Suppression of Two Different Stop Codons. Biotechnol. Bioeng. 2017, 114 (5), 1065–1073. 10.1002/bit.26226. [DOI] [PubMed] [Google Scholar]
- Albayrak C.; Swartz J. R. Cell-Free Co-Production of an Orthogonal Transfer RNA Activates Efficient Site-Specific Non-Natural Amino Acid Incorporation. Nucleic Acids Res. 2013, 41 (11), 5949–5963. 10.1093/nar/gkt226. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee K. B.; Hou C. Y.; Kim C. E.; Kim D. M.; Suga H.; Kang T. J. Genetic Code Expansion by Degeneracy Reprogramming of Arginyl Codons. Chembiochem 2016, 17, 1198–1201. 10.1002/cbic.201600111. [DOI] [PubMed] [Google Scholar]
- Salehi A. S. M.; Smith M. T.; Schinn S. M.; Hunt J. M.; Muhlestein C.; Diray-Arce J.; Nielsen B. L.; Bundy B. C. Efficient TRNA Degradation and Quantification in Escherichia Coli Cell Extract Using RNase-Coated Magnetic Beads: A Key Step toward Codon Emancipation. Biotechnol. Prog. 2017, 33 (5), 1401–1407. 10.1002/btpr.2511. [DOI] [PubMed] [Google Scholar]
- Cui Z.; Stein V.; Tnimov Z.; Mureev S.; Alexandrov K. Semisynthetic TRNA Complement Mediates in Vitro Protein Synthesis. J. Am. Chem. Soc. 2015, 137 (13), 4404–4413. 10.1021/ja5131963. [DOI] [PubMed] [Google Scholar]
- Cui Z.; Mureev S.; Polinkovsky M. E.; Tnimov Z.; Guo Z.; Durek T.; Jones A.; Alexandrov K. Combining Sense and Nonsense Codon Reassignment for Site-Selective Protein Modification with Unnatural Amino Acids. ACS Synth. Biol. 2017, 6 (3), 535–544. 10.1021/acssynbio.6b00245. [DOI] [PubMed] [Google Scholar]
- Cui Z.; Wu Y.; Mureev S.; Alexandrov K. Oligonucleotide-Mediated TRNA Sequestration Enables One-Pot Sense Codon Reassignment in Vitro. Nucleic Acids Res. 2018, 46 (12), 6387–6400. 10.1093/nar/gky365. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Goerke A. R.; Swartz J. R. High-Level Cell-Free Synthesis Yields of Proteins Containing Site-Specific Non-Natural Amino Acids. Biotechnol. Bioeng. 2009, 102 (2), 400–416. 10.1002/bit.22070. [DOI] [PubMed] [Google Scholar]
- Ozawa K.; Loscha K. V.; Kuppan K. V.; Loh C. T.; Dixon N. E.; Otting G. High-Yield Cell-Free Protein Synthesis for Site-Specific Incorporation of Unnatural Amino Acids at Two Sites. Biochem. Biophys. Res. Commun. 2012, 418 (4), 652–656. 10.1016/j.bbrc.2012.01.069. [DOI] [PubMed] [Google Scholar]
- Hong S. H.; Ntai I.; Haimovich A. D.; Kelleher N. L.; Isaacs F. J.; Jewett M. C. Cell-Free Protein Synthesis from a Release Factor 1 Deficient Escherichia Coli Activates Efficient and Multiple Site-Specific Nonstandard Amino Acid Incorporation. ACS Synth. Biol. 2014, 3 (6), 398–409. 10.1021/sb400140t. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hou J.; Chen X.; Jiang N.; Wang Y.; Cui Y.; Ma L.; Lin Y.; Lu Y. Toward Efficient Multiple-Site Incorporation of Unnatural Amino Acids Using Cell-Free Translation System. Synth. Syst. Biotechnol. 2022, 7 (1), 522–532. 10.1016/j.synbio.2021.12.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Seki E.; Yanagisawa T.; Kuratani M.; Sakamoto K.; Yokoyama S. Fully Productive Cell-Free Genetic Code Expansion by Structure- Based Engineering of Methanomethylophilus Alvus Pyrrolysyl-TRNA Synthetase. ACS Synth. Biol. 2020, 9, 718. 10.1021/acssynbio.9b00288. [DOI] [PubMed] [Google Scholar]
- Gerrits M.; Budisa N.; Merk H. Site-Specific Chemoselective Pyrrolysine Analogues Incorporation Using the Cell-Free Protein Synthesis System. ACS Synth. Biol. 2019, 8, 381–390. 10.1021/acssynbio.8b00421. [DOI] [PubMed] [Google Scholar]
- Mittelstaet J.; Konevega A. L.; Rodnina M. V. A Kinetic Safety Gate Controlling the Delivery of Unnatural Amino Acids to the Ribosome. J. Am. Chem. Soc. 2013, 135 (45), 17031–17038. 10.1021/ja407511q. [DOI] [PubMed] [Google Scholar]
- Doi Y.; Ohtsuki T.; Shimizu Y.; Ueda T.; Sisido M. Elongation Factor Tu Mutants Expand Amino Acid Tolerance of Protein Biosynthesis System. J. Am. Chem. Soc. 2007, 129 (46), 14458–14462. 10.1021/ja075557u. [DOI] [PubMed] [Google Scholar]
- Ohtsuki T.; Yamamoto H.; Doi Y.; Sisido M. Use of EF-Tu Mutants for Determining and Improving Aminoacylation Efficiency and for Purifying Aminoacyl TRNAs with Non-Natural Amino Acids. J. Biochem. 2010, 148 (2), 239–246. 10.1093/jb/mvq053. [DOI] [PubMed] [Google Scholar]
- Gan R.; Perez J. G.; Carlson E. D.; Ntai I.; Isaacs F. J.; Kelleher N. L.; Jewett M. C. Translation System Engineering in Escherichia Coli Enhances Non-Canonical Amino Acid Incorporation into Proteins. Biotechnol. Bioeng. 2017, 114 (5), 1074–1086. 10.1002/bit.26239. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yin G.; Stephenson H. T.; Yang J.; Li X.; Armstrong S. M.; Heibeck T. H.; Tran C.; Masikat M. R.; Zhou S.; Stafford R. L.; Yam A. Y.; Lee J.; Steiner A. R.; Gill A.; Penta K.; Pollitt S.; Baliga R.; Murray C. J.; Thanos C. D.; McEvoy L. M.; Sato A. K.; Hallam T. J. RF1 Attenuation Enables Efficient Non-Natural Amino Acid Incorporation for Production of Homogeneous Antibody Drug Conjugates. Sci. Rep. 2017, 7 (1), 1–13. 10.1038/s41598-017-03192-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee J.; Schwarz K. J.; Kim D. S.; Moore J. S.; Jewett M. C. Ribosome-Mediated Polymerization of Long Chain Carbon and Cyclic Amino Acids into Peptides in Vitro. Nat. Commun. 2020, 11 (1), 1–8. 10.1038/s41467-020-18001-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Choy H. E. Regulated Transcription in a Complete Ribosome-Free in Vitro System of Escherichia Coli. Methods Enzymol. 1996, 274, 3–8. 10.1016/S0076-6879(96)74003-X. [DOI] [PubMed] [Google Scholar]
- Liu Y.; Davis R. G.; Thomas P. M.; Kelleher N. L.; Jewett M. C. In Vitro-Constructed Ribosomes Enable Multi-Site Incorporation of Noncanonical Amino Acids into Proteins. Biochemistry 2021, 60 (3), 161–169. 10.1021/acs.biochem.0c00829. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Malde A. K.; Hill T. A.; Iyer A.; Fairlie D. P. Crystal Structures of Protein-Bound Cyclic Peptides. Chem. Rev. 2019, 119 (17), 9861–9914. 10.1021/acs.chemrev.8b00807. [DOI] [PubMed] [Google Scholar]
- Morimoto J.; Hayashi Y.; Suga H. Discovery of Macrocyclic Peptides Armed with a Mechanism-Based Warhead: Isoform-Selective Inhibition of Human Deacetylase SIRT2. Angew. Chem., Int. Ed. Engl. 2012, 51 (14), 3423–3427. 10.1002/anie.201108118. [DOI] [PubMed] [Google Scholar]
- Yin Y.; Ochi N.; Craven T. W.; Baker D.; Takigawa N.; Suga H. De Novo Carborane-Containing Macrocyclic Peptides Targeting Human Epidermal Growth Factor Receptor. J. Am. Chem. Soc. 2019, 141 (49), 19193–19197. 10.1021/jacs.9b09106. [DOI] [PubMed] [Google Scholar]
- Passioura T.; Suga H. A RaPID Way to Discover Nonstandard Macrocyclic Peptide Modulators of Drug Targets. Chem. Commun. 2017, 53 (12), 1931–1940. 10.1039/C6CC06951G. [DOI] [PubMed] [Google Scholar]
- Goto Y.; Suga H. The RaPID Platform for the Discovery of Pseudo-Natural Macrocyclic Peptides. Acc. Chem. Res. 2021, 54 (18), 3604–3617. 10.1021/acs.accounts.1c00391. [DOI] [PubMed] [Google Scholar]
- Goto Y.; Suga H. Engineering of RiPP Pathways for the Production of Artificial Peptides Bearing Various Non-Proteinogenic Structures. Curr. Opin. Chem. Biol. 2018, 46, 82–90. 10.1016/j.cbpa.2018.06.014. [DOI] [PubMed] [Google Scholar]
- Tsutsumi H.; Kuroda T.; Kimura H.; Goto Y.; Suga H. Posttranslational Chemical Installation of Azoles into Translated Peptides. Nat. Commun. 2021, 12 (1), 696–696. 10.1038/s41467-021-20992-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fleming S. R.; Bartges T. E.; Vinogradov A. A.; Kirkpatrick C. L.; Goto Y.; Suga H.; Hicks L. M.; Bowers A. A. Flexizyme-Enabled Benchtop Biosynthesis of Thiopeptides. J. Am. Chem. Soc. 2019, 141 (2), 758–762. 10.1021/jacs.8b11521. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Iwane Y.; Hitomi A.; Murakami H.; Katoh T.; Goto Y.; Suga H. Expanding the Amino Acid Repertoire of Ribosomal Polypeptide Synthesis via the Artificial Division of Codon Boxes. Nat. Chem. 2016, 8 (4), 317–325. 10.1038/nchem.2446. [DOI] [PubMed] [Google Scholar]
- Hibi K.; Amikura K.; Sugiura N.; Masuda K.; Ohno S.; Yokogawa T.; Ueda T.; Shimizu Y. Reconstituted Cell-Free Protein Synthesis Using in Vitro Transcribed TRNAs. Communications Biology 2020, 3 (1), 350–350. 10.1038/s42003-020-1074-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fujino T.; Tozaki M.; Murakami H. An Amino Acid-Swapped Genetic Code. ACS Synth. Biol. 2020, 9, 2703. 10.1021/acssynbio.0c00196. [DOI] [PubMed] [Google Scholar]
- Calles J.; Justice I.; Brinkley D.; Garcia A.; Endy D. Fail-Safe Genetic Codes Designed to Intrinsically Contain Engineered Organisms. Nucleic Acids Res. 2019, 47 (19), 10439–10451. 10.1093/nar/gkz745. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Terasaka N.; Hayashi G.; Katoh T.; Suga H. An Orthogonal Ribosome-TRNA Pair via Engineering of the Peptidyl Transferase Center. Nat. Chem. Biol. 2014, 10 (7), 555–557. 10.1038/nchembio.1549. [DOI] [PubMed] [Google Scholar]
- Short G. F. 3rd; Golovine S. Y.; Hecht S. M. Effects of Release Factor 1 on in Vitro Protein Translation and the Elaboration of Proteins Containing Unnatural Amino Acids. Biochemistry 1999, 38 (27), 8808–8819. 10.1021/bi990281r. [DOI] [PubMed] [Google Scholar]
- Lajoie M. J.; Rovner A. J.; Goodman D. B.; Aerni H.-R.; Haimovich A. D.; Kuznetsov G.; Mercer J. a.; Wang H. H.; Carr P. a.; Mosberg J. a.; Rohland N.; Schultz P. G.; Jacobson J. M.; Rinehart J.; Church G. M.; Isaacs F. J. Genomically Recoded Organisms Expand Biological Functions. Science 2013, 342, 357–360. 10.1126/science.1241459. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mukai T.; Hoshi H.; Ohtake K.; Takahashi M.; Yamaguchi A.; Hayashi A.; Yokoyama S.; Sakamoto K. Highly Reproductive Escherichia Coli Cells with No Specific Assignment to the UAG Codon. Sci. Rep. 2015, 5, 1–9. 10.1038/srep09699. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Seki E.; Yanagisawa T.; Yokoyama S. Cell-Free Protein Synthesis for Multiple Site-Specific Incorporation of Noncanonical Amino Acids Using Cell Extracts from RF-1 Deletion E. Coli Strains. Methods Mol. Biol. 2018, 1728, 49–65. 10.1007/978-1-4939-7574-7_3. [DOI] [PubMed] [Google Scholar]
- Seki K.; Galindo J. L.; Karim A. S.; Jewett M. C. A Cell-Free Gene Expression Platform for Discovering and Characterizing Stop Codon Suppressing TRNAs. ACS Chem. Biol. 2023, 18 (6), 1324–1334. 10.1021/acschembio.3c00051. [DOI] [PubMed] [Google Scholar]
- Singh-Blom A.; Hughes R. A.; Ellington A. D. An Amino Acid Depleted Cell-Free Protein Synthesis System for the Incorporation of Non-Canonical Amino Acid Analogs into Proteins. J. Biotechnol. 2014, 178 (1), 12–22. 10.1016/j.jbiotec.2014.02.009. [DOI] [PubMed] [Google Scholar]
- Worst E. G.; Exner M. P.; De Simone A.; Schenkelberger M.; Noireaux V.; Budisa N.; Ott A.. Residue-Specific Incorporation of Noncanonical Amino Acids into Model Proteins Using an Escherichia Coli Cell-Free Transcription-Translation System. J. Vis. Exp. 2016, No. (114), . 10.3791/54273. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kigawa T.; Muto Y.; Yokoyama S. Cell-Free Synthesis and Amino Acid-Selective Stable Isotope Labeling of Proteins for NMR Analysis. J. Biomol. NMR 1995, 6 (2), 129–134. 10.1007/BF00211776. [DOI] [PubMed] [Google Scholar]
- Zimmerman E. S.; Heibeck T. H.; Gill A.; Li X.; Murray C. J.; Madlansacay M. R.; Tran C.; Uter N. T.; Yin G.; Rivers P. J.; Yam A. Y.; Wang W. D.; Steiner A. R.; Bajad S. U.; Penta K.; Yang W.; Hallam T. J.; Thanos C. D.; Sato A. K. Production of Site-Specific Antibody-Drug Conjugates Using Optimized Non-Natural Amino Acids in a Cell-Free Expression System. Bioconjugate Chem. 2014, 25 (2), 351–361. 10.1021/bc400490z. [DOI] [PubMed] [Google Scholar]
- Ugwumba I. N.; Ozawa K.; Xu Z. Q.; Ely F.; Foo J. L.; Herlt A. J.; Coppin C.; Brown S.; Taylor M. C.; Ollis D. L.; Mander L. N.; Schenk G.; Dixon N. E.; Otting G.; Oakeshott J. G.; Jackson C. J. Improving a Natural Enzyme Activity through Incorporation of Unnatural Amino Acids. J. Am. Chem. Soc. 2011, 133 (2), 326–333. 10.1021/ja106416g. [DOI] [PubMed] [Google Scholar]
- Zubi Y. S.; Seki K.; Li Y.; Hunt A. C.; Liu B.; Roux B.; Jewett M. C.; Lewis J. C. Metal-Responsive Regulation of Enzyme Catalysis Using Genetically Encoded Chemical Switches. Nat. Commun. 2022, 13 (1), 1864. 10.1038/s41467-022-29239-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu W. R.; Wang Y.-S.; Wan W. Synthesis of Proteins with Defined Posttranslational Modifications Using the Genetic Noncanonical Amino Acid Incorporation Approach. Mol. Biosyst. 2011, 7 (1), 38–47. 10.1039/C0MB00216J. [DOI] [PubMed] [Google Scholar]
- Loscha K. V.; Herlt A. J.; Qi R.; Huber T.; Ozawa K.; Otting G. Multiple-Site Labeling of Proteins with Unnatural Amino Acids. Angewandte Chemie - International Edition 2012, 51 (9), 2243–2246. 10.1002/anie.201108275. [DOI] [PubMed] [Google Scholar]
- Nguyen T. H. D.; Ozawa K.; Stanton-Cook M.; Barrow R.; Huber T.; Otting G. Generation of Pseudocontact Shifts in Protein NMR Spectra with a Genetically Encoded Cobalt(II)-Binding Amino Acid. Angew. Chem., Int. Ed. Engl. 2011, 50 (3), 692–694. 10.1002/anie.201005672. [DOI] [PubMed] [Google Scholar]
- Taki M.; Hohsaka T.; Murakami H.; Taira K.; Sisido M. A Non-Natural Amino Acid for Efficient Incorporation into Proteins as a Sensitive Fluorescent Probe. FEBS Lett. 2001, 507 (1), 35–38. 10.1016/S0014-5793(01)02935-0. [DOI] [PubMed] [Google Scholar]
- Ugwumba I. N.; Ozawa K.; de la Cruz L.; Xu Z.-Q.; Herlt A. J.; Hadler K. S.; Coppin C.; Brown S. E.; Schenk G.; Oakeshott J. G.; Otting G. Using a Genetically Encoded Fluorescent Amino Acid as a Site-Specific Probe to Detect Binding of Low-Molecular-Weight Compounds. Assay Drug Dev. Technol. 2011, 9 (1), 50–57. 10.1089/adt.2010.0306. [DOI] [PubMed] [Google Scholar]
- Kajihara D.; Abe R.; Iijima I.; Komiyama C.; Sisido M.; Hohsaka T. FRET Analysis of Protein Conformational Change through Position-Specific Incorporation of Fluorescent Amino Acids. Nat. Methods 2006, 3 (11), 923–929. 10.1038/nmeth945. [DOI] [PubMed] [Google Scholar]
- Anderson R. D. 3rd; Zhou J.; Hecht S. M. Fluorescence Resonance Energy Transfer between Unnatural Amino Acids in a Structurally Modified Dihydrofolate Reductase. J. Am. Chem. Soc. 2002, 124 (33), 9674–9675. 10.1021/ja0205939. [DOI] [PubMed] [Google Scholar]
- Taki M.; Hohsaka T.; Murakami H.; Taira K.; Sisido M. Position-Specific Incorporation of a Fluorophore - Quencher Pair into a Single Streptavidin through Orthogonal Four-Base Codon/Anticodon Pairs. J. Am. Chem. Soc. 2002, 124 (49), 14586–14590. 10.1021/ja017714+. [DOI] [PubMed] [Google Scholar]
- Adiga R.; Al-Adhami M.; Andar A.; Borhani S.; Brown S.; Burgenson D.; Cooper M. A.; Deldari S.; Frey D. D.; Ge X.; Guo H.; Gurramkonda C.; Jensen P.; Kostov Y.; LaCourse W.; Liu Y.; Moreira A.; Mupparapu K.; Peñalber-Johnstone C.; Pilli M.; Punshon-Smith B.; Rao A.; Rao G.; Rauniyar P.; Snovida S.; Taurani K.; Tilahun D.; Tolosa L.; Tolosa M.; Tran K.; Vattem K.; Veeraraghavan S.; Wagner B.; Wilhide J.; Wood D. W.; Zuber A. Point-of-Care Production of Therapeutic Proteins of Good-Manufacturing-Practice Quality. Nat. Biomed Eng. 2018, 2 (9), 675–686. 10.1038/s41551-018-0259-1. [DOI] [PubMed] [Google Scholar]
- Sullivan C. J.; Pendleton E. D.; Sasmor H. H.; Hicks W. L.; Farnum J. B.; Muto M.; Amendt E. M.; Schoborg J. A.; Martin R. W.; Clark L. G.; Anderson M. J.; Choudhury A.; Fior R.; Lo Y.-H.; Griffey R. H.; Chappell S. A.; Jewett M. C.; Mauro V. P.; Dresios J. A Cell-free Expression and Purification Process for Rapid Production of Protein Biologics. Biotechnol. J. 2016, 11 (2), 238–248. 10.1002/biot.201500214. [DOI] [PubMed] [Google Scholar]
- Yin G.; Garces E. D.; Yang J.; Zhang J.; Tran C.; Steiner A. R.; Roos C.; Bajad S.; Hudak S.; Penta K.; Zawada J.; Pollitt S.; Murray C. J. Aglycosylated Antibodies and Antibody Fragments Produced in a Scalable in Vitro Transcription-Translation System. MAbs 2012, 4 (2), 217–225. 10.4161/mabs.4.2.19202. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Voloshin A. M.; Swartz J. R.. Large-Scale Batch Reactions for Cell-Free Protein Synthesis. In Cell-Free Protein Synthesis; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2008; pp 207–235. 10.1002/9783527622702.ch12. [DOI] [Google Scholar]
- Rasor B. J.; Vögeli B.; Landwehr G. M.; Bogart J. W.; Karim A. S.; Jewett M. C. Toward Sustainable, Cell-Free Biomanufacturing. Curr. Opin. Biotechnol. 2021, 69, 136–144. 10.1016/j.copbio.2020.12.012. [DOI] [PubMed] [Google Scholar]
- You C.; Shi T.; Li Y.; Han P.; Zhou X.; Zhang Y.-H. P. An in Vitro Synthetic Biology Platform for the Industrial Biomanufacturing of Myo-Inositol from Starch. Biotechnol. Bioeng. 2017, 114 (8), 1855–1864. 10.1002/bit.26314. [DOI] [PubMed] [Google Scholar]
- Calhoun K. A.; Swartz J. R. Total Amino Acid Stabilization during Cell-Free Protein Synthesis Reactions. J. Biotechnol. 2006, 123 (2), 193–203. 10.1016/j.jbiotec.2005.11.011. [DOI] [PubMed] [Google Scholar]
- Hunt J. P.; Yang S. O.; Wilding K. M.; Bundy B. C. The Growing Impact of Lyophilized Cell-Free Protein Expression Systems. Bioengineered 2017, 8 (4), 325–330. 10.1080/21655979.2016.1241925. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McNerney M. P.; Zhang Y.; Steppe P.; Silverman A. D.; Jewett M. C.; Styczynski M. P. Point-of-Care Biomarker Quantification Enabled by Sample-Specific Calibration. Sci. Adv. 2019, 5 (9), eaax4473 10.1126/sciadv.aax4473. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wilding K. M.; Zhao E. L.; Earl C. C.; Bundy B. C. Thermostable Lyoprotectant-Enhanced Cell-Free Protein Synthesis for on-Demand Endotoxin-Free Therapeutic Production. N. Biotechnol. 2019, 53, 73–80. 10.1016/j.nbt.2019.07.004. [DOI] [PubMed] [Google Scholar]
- Yang J.; Cui Y.; Cao Z.; Ma S.; Lu Y. Strategy Exploration for Developing Robust Lyophilized Cell-Free Systems. Biotechnology Notes 2021, 2, 44–50. 10.1016/j.biotno.2021.08.004. [DOI] [Google Scholar]
- Williams A. J.; Warfel K. F.; Desai P.; Li J.; Lee J.-J.; Wong D. A.; Nguyen P. M.; Qin Y.; Sobol S. E.; Jewett M. C.; Chang Y.-F.; DeLisa M. P.. A Low-Cost Recombinant Glycoconjugate Vaccine Confers Immunogenicity and Protection against Enterotoxigenic Escherichia Coli Infections in Mice. Front. Mol. Biosci. 2023, 10. 10.3389/fmolb.2023.1085887. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stamatis C.; Farid S. S. Process Economics Evaluation of Cell-free Synthesis for the Commercial Manufacture of Antibody Drug Conjugates. Biotechnol. J. 2021, 16, 2000238. 10.1002/biot.202000238. [DOI] [PubMed] [Google Scholar]
- Sierecki E.; Giles N.; Polinkovsky M.; Moustaqil M.; Alexandrov K.; Gambin Y. A Cell-Free Approach to Accelerate the Study of Protein-Protein Interactions in Vitro. Interface Focus 2013, 3 (5), 20130018. 10.1098/rsfs.2013.0018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wiens M. D.; Campbell R. E. Surveying the Landscape of Optogenetic Methods for Detection of Protein-Protein Interactions. Wiley Interdiscip. Rev. Syst. Biol. Med. 2018, 10 (3), 1–15. 10.1002/wsbm.1415. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Porter J. R.; Stains C. I.; Jester B. W.; Ghosh I. A General and Rapid Cell-Free Approach for the Interrogation of Protein-Protein, Protein-DNA, and Protein-RNA Interactions and Their Antagonists Utilizing Split-Protein Reporters. J. Am. Chem. Soc. 2008, 130 (20), 6488–6497. 10.1021/ja7114579. [DOI] [PubMed] [Google Scholar]
- Campbell S. T.; Carlson K. J.; Buchholz C. J.; Helmers M. R.; Ghosh I. Mapping the BH3 Binding Interface of Bcl-x L, Bcl-2, and Mcl-1 Using Split-Luciferase Reassembly. Biochemistry 2015, 54 (16), 2632–2643. 10.1021/bi501505y. [DOI] [PubMed] [Google Scholar]
- Henchey L. K.; Porter J. R.; Ghosh I.; Arora P. S. High Specificity in Protein Recognition by Hydrogen-Bond-Surrogate α-Helices: Selective Inhibition of the P53/MDM2 Complex. Chembiochem 2010, 11 (15), 2104–2107. 10.1002/cbic.201000378. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Porter J. R.; Helmers M. R.; Wang P.; Furman J. L.; Joy S. T.; Arora P. S.; Ghosh I. Profiling Small Molecule Inhibitors against Helix-Receptor Interactions: The Bcl-2 Family Inhibitor BH3I-1 Potently Inhibits P53/HDM2. Chem. Commun. 2010, 46 (42), 8020–8022. 10.1039/c0cc02969f. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Doi N.; Takashima H.; Kinjo M.; Sakata K.; Kawahashi Y.; Oishi Y.; Oyama R.; Miyamoto-Sato E.; Sawasaki T.; Endo Y.; Yanagawa H. Novel Fluorescence Labeling and High-Throughput Assay Technologies for in Vitro Analysis of Protein Interactions. Genome Res. 2002, 12 (3), 487–492. 10.1101/gr.218802. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sierecki E.; Stevers L. M.; Giles N.; Polinkovsky M. E.; Moustaqil M.; Mureev S.; Johnston W. A.; Dahmer-Heath M.; Skalamera D.; Gonda T. J.; Gabrielli B.; Collins B. M.; Alexandrov K.; Gambin Y. Rapid Mapping of Interactions between Human SNX-BAR Proteins Measured In Vitro by AlphaScreen and Single-Molecule Spectroscopy. Mol. Cell. Proteomics 2014, 13 (9), 2233–2245. 10.1074/mcp.M113.037275. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gambin Y.; Ariotti N.; McMahon K.-A.; Bastiani M.; Sierecki E.; Kovtun O.; Polinkovsky M. E; Magenau A.; Jung W.; Okano S.; Zhou Y.; Leneva N.; Mureev S.; Johnston W.; Gaus K.; Hancock J. F; Collins B. M; Alexandrov K.; Parton R. G Single-Molecule Analysis Reveals Self Assembly and Nanoscale Segregation of Two Distinct Cavin Subcomplexes on Caveolae. Elife 2014, 3, e01434 10.7554/eLife.01434. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Han S. P.; Gambin Y.; Gomez G. A.; Verma S.; Giles N.; Michael M.; Wu S. K.; Guo Z.; Johnston W.; Sierecki E.; Parton R. G.; Alexandrov K.; Yap A. S. Cortactin Scaffolds Arp2/3 and WAVE2 at the Epithelial Zonula Adherens. J. Biol. Chem. 2014, 289 (11), 7764–7775. 10.1074/jbc.M113.544478. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aydin S. A Short History, Principles, and Types of ELISA, and Our Laboratory Experience with Peptide/Protein Analyses Using ELISA. Peptides 2015, 72, 4–15. 10.1016/j.peptides.2015.04.012. [DOI] [PubMed] [Google Scholar]
- Ojima-Kato T.; Morishita S.; Uchida Y.; Nagai S.; Kojima T.; Nakano H. Rapid Generation of Monoclonal Antibodies from Single B Cells by Ecobody Technology. Antibodies 2018, 7 (4), 38. 10.3390/antib7040038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ojima-Kato T.; Hashimura D.; Kojima T.; Minabe S.; Nakano H. In Vitro Generation of Rabbit Anti-Listeria Monocytogenes Monoclonal Antibody Using Single Cell Based RT-PCR Linked Cell-Free Expression Systems. J. Immunol. Methods 2015, 427, 58–65. 10.1016/j.jim.2015.10.001. [DOI] [PubMed] [Google Scholar]
- Beaudet L.; Rodriguez-Suarez R.; Venne M.-H.; Caron M.; Bédard J.; Brechler V.; Parent S.; Bielefeld-Sévigny M. AlphaLISA Immunoassays: The No-Wash Alternative to ELISAs for Research and Drug Discovery. Nat. Methods 2008, 5 (12), an8–an9. 10.1038/nmeth.f.230. [DOI] [Google Scholar]
- Matsuoka K.; Komori H.; Nose M.; Endo Y.; Sawasaki T. Simple Screening Method for Autoantigen Proteins Using the N-Terminal Biotinylated Protein Library Produced by Wheat Cell-Free Synthesis. J. Proteome Res. 2010, 9 (8), 4264–4273. 10.1021/pr9010553. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhou Y.; Asahara H.; Schneider N.; Dranchak P.; Inglese J.; Chong S. Engineering Bacterial Transcription Regulation to Create a Synthetic in Vitro Two-Hybrid System for Protein Interaction Assays. J. Am. Chem. Soc. 2014, 136 (40), 14031–14038. 10.1021/ja502512g. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ceroni A.; Sibani S.; Baiker A.; Pothineni V. R.; Bailer S. M.; LaBaer J.; Haas J.; Campbell C. J. Systematic Analysis of the IgG Antibody Immune Response against Varicella Zoster Virus (VZV) Using a Self-Assembled Protein Microarray. Mol. Biosyst. 2010, 6 (9), 1604–1610. 10.1039/c003798b. [DOI] [PubMed] [Google Scholar]
- Wang J.; Barker K.; Steel J.; Park J.; Saul J.; Festa F.; Wallstrom G.; Yu X.; Bian X.; Anderson K. S.; Figueroa J. D.; LaBaer J.; Qiu J. A Versatile Protein Microarray Platform Enabling Antibody Profiling against Denatured Proteins. Proteomics Clin. Appl. 2013, 7 (5–6), 378–383. 10.1002/prca.201200062. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sibani S.; LaBaer J. Immunoprofiling Using NAPPA Protein Microarrays. Methods Mol. Biol. 2011, 723, 149–161. 10.1007/978-1-61779-043-0_10. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Díez P.; González-González M.; Lourido L.; Dégano R. M.; Ibarrola N.; Casado-Vela J.; LaBaer J.; Fuentes M. NAPPA as a Real New Method for Protein Microarray Generation. Microarrays (Basel) 2015, 4 (2), 214–227. 10.3390/microarrays4020214. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stoevesandt O.; Taussig M. J.; He M. Protein Microarrays: High-Throughput Tools for Proteomics. Expert Rev. Proteomics 2009, 6 (2), 145–157. 10.1586/epr.09.2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Garcia-Cordero J. L.; Maerkl S. J. Mechanically Induced Trapping of Molecular Interactions and Its Applications. J. Lab. Autom. 2016, 21 (3), 356–367. 10.1177/2211068215578586. [DOI] [PubMed] [Google Scholar]
- Cui N.; Zhang H.; Schneider N.; Tao Y.; Asahara H.; Sun Z.; Cai Y.; Koehler S. A.; de Greef T. F. A.; Abbaspourrad A.; Weitz D. A.; Chong S. A Mix-and-Read Drop-Based in Vitro Two-Hybrid Method for Screening High-Affinity Peptide Binders. Sci. Rep. 2016, 6 (1), 22575. 10.1038/srep22575. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Linciano S.; Pluda S.; Bacchin A.; Angelini A. Molecular Evolution of Peptides by Yeast Surface Display Technology. Medchemcomm 2019, 10 (9), 1569–1580. 10.1039/C9MD00252A. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Plückthun A. Ribosome Display: A Perspective. Methods Mol. Biol. 2012, 805, 3–28. 10.1007/978-1-61779-379-0_1. [DOI] [PubMed] [Google Scholar]
- Lipovsek D.; Plückthun A. In-Vitro Protein Evolution by Ribosome Display and MRNA Display. J. Immunol. Methods 2004, 290 (1–2), 51–67. 10.1016/j.jim.2004.04.008. [DOI] [PubMed] [Google Scholar]
- Zahnd C.; Amstutz P.; Plückthun A. Ribosome Display: Selecting and Evolving Proteins in Vitro That Specifically Bind to a Target. Nat. Methods 2007, 4 (3), 269–279. 10.1038/nmeth1003. [DOI] [PubMed] [Google Scholar]
- Luginbühl B.; Kanyo Z.; Jones R. M.; Fletterick R. J.; Prusiner S. B.; Cohen F. E.; Williamson R. A.; Burton D. R.; Plückthun A. Directed Evolution of an Anti-Prion Protein ScFv Fragment to an Affinity of 1 PM and Its Structural Interpretation. J. Mol. Biol. 2006, 363 (1), 75–97. 10.1016/j.jmb.2006.07.027. [DOI] [PubMed] [Google Scholar]
- Hanes J.; Schaffitzel C.; Knappik A.; Plückthun A. Picomolar Affinity Antibodies from a Fully Synthetic Naive Library Selected and Evolved by Ribosome Display. Nat. Biotechnol. 2000, 18 (12), 1287–1292. 10.1038/82407. [DOI] [PubMed] [Google Scholar]
- Stafford R. L.; Matsumoto M. L.; Yin G.; Cai Q.; Fung J. J.; Stephenson H.; Gill A.; You M.; Lin S.-H.; Wang W. D.; Masikat M. R.; Li X.; Penta K.; Steiner A. R.; Baliga R.; Murray C. J.; Thanos C. D.; Hallam T. J.; Sato A. K. In Vitro Fab Display: A Cell-Free System for IgG Discovery. Protein Eng. Des. Sel. 2014, 27 (4), 97–109. 10.1093/protein/gzu002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Binz H. K.; Amstutz P.; Kohl A.; Stumpp M. T.; Briand C.; Forrer P.; Grütter M. G.; Plückthun A. High-Affinity Binders Selected from Designed Ankyrin Repeat Protein Libraries. Nat. Biotechnol. 2004, 22 (5), 575–582. 10.1038/nbt962. [DOI] [PubMed] [Google Scholar]
- Zimmermann I.; Egloff P.; Hutter C. A. J.; Arnold F. M.; Stohler P.; Bocquet N.; Hug M. N.; Huber S.; Siegrist M.; Hetemann L.; Gera J.; Gmür S.; Spies P.; Gygax D.; Geertsma E. R.; Dawson R. J. P.; Seeger M. A.. Synthetic Single Domain Antibodies for the Conformational Trapping of Membrane Proteins. Elife 2018, 7. 10.7554/eLife.34317. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Custódio T. F.; Das H.; Sheward D. J.; Hanke L.; Pazicky S.; Pieprzyk J.; Sorgenfrei M.; Schroer M. A.; Gruzinov A. Y.; Jeffries C. M.; Graewert M. A.; Svergun D. I.; Dobrev N.; Remans K.; Seeger M. A.; McInerney G. M.; Murrell B.; Hällberg B. M.; Löw C. Selection, Biophysical and Structural Analysis of Synthetic Nanobodies That Effectively Neutralize SARS-CoV-2. Nat. Commun. 2020, 11 (1), 5588. 10.1038/s41467-020-19204-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zimmermann I.; Egloff P.; Hutter C. A. J.; Kuhn B. T.; Bräuer P.; Newstead S.; Dawson R. J. P.; Geertsma E. R.; Seeger M. A. Generation of Synthetic Nanobodies against Delicate Proteins. Nat. Protoc. 2020, 15, 1707. 10.1038/s41596-020-0304-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen X.; Gentili M.; Hacohen N.; Regev A. A Cell-Free Nanobody Engineering Platform Rapidly Generates SARS-CoV-2 Neutralizing Nanobodies. Nat. Commun. 2021, 12 (1), 5506. 10.1038/s41467-021-25777-z. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Veesler D.; Dreier B.; Blangy S.; Lichière J.; Tremblay D.; Moineau S.; Spinelli S.; Tegoni M.; Plückthun A.; Campanacci V.; Cambillau C. Crystal Structure and Function of a DARPin Neutralizing Inhibitor of Lactococcal Phage TP901–1: Comparison of DARPin and Camelid VHH Binding Mode. J. Biol. Chem. 2009, 284 (44), 30718–30726. 10.1074/jbc.M109.037812. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zahnd C.; Wyler E.; Schwenk J. M.; Steiner D.; Lawrence M. C.; McKern N. M.; Pecorari F.; Ward C. W.; Joos T. O.; Plückthun A. A Designed Ankyrin Repeat Protein Evolved to Picomolar Affinity to Her2. J. Mol. Biol. 2007, 369 (4), 1015–1028. 10.1016/j.jmb.2007.03.028. [DOI] [PubMed] [Google Scholar]
- Matsuura T.; Yanagida H.; Ushioda J.; Urabe I.; Yomo T. Nascent Chain, MRNA, and Ribosome Complexes Generated by a Pure Translation System. Biochem. Biophys. Res. Commun. 2007, 352 (2), 372–377. 10.1016/j.bbrc.2006.11.050. [DOI] [PubMed] [Google Scholar]
- Galán A.; Comor L.; Horvatić A.; Kuleš J.; Guillemin N.; Mrljak V.; Bhide M. Library-Based Display Technologies: Where Do We Stand?. Mol. Biosyst. 2016, 12 (8), 2342–2358. 10.1039/C6MB00219F. [DOI] [PubMed] [Google Scholar]
- In Vitro Virus: Bonding of MRNA Bearing Puromycin at the 3′-Terminal End to the C-Terminal End of Its Encoded Protein on the Ribosome in Vitro. FEBS Lett. 1997, 414 ( (2), ), 405–408. 10.1016/S0014-5793(97)01026-0. [DOI] [PubMed] [Google Scholar]
- Wilson D. S.; Keefe A. D.; Szostak J. W. The Use of MRNA Display to Select High-Affinity Protein-Binding Peptides. Proc. Natl. Acad. Sci. U. S. A. 2001, 98 (7), 3750–3755. 10.1073/pnas.061028198. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Xu L.; Aha P.; Gu K.; Kuimelis R. G.; Kurz M.; Lam T.; Lim A. C.; Liu H.; Lohse P. A.; Sun L.; Weng S.; Wagner R. W.; Lipovsek D. Directed Evolution of High-Affinity Antibody Mimics Using MRNA Display. Chem. Biol. 2002, 9 (8), 933–942. 10.1016/S1074-5521(02)00187-4. [DOI] [PubMed] [Google Scholar]
- Suzuki T.; Mochizuki Y.; Kimura S.; Akazawa-Ogawa Y.; Hagihara Y.; Nemoto N. Anti-Survivin Single-Domain Antibodies Derived from an Artificial Library Including Three Synthetic Random Regions by in Vitro Selection Using CDNA Display. Biochem. Biophys. Res. Commun. 2018, 503 (3), 2054–2060. 10.1016/j.bbrc.2018.07.158. [DOI] [PubMed] [Google Scholar]
- Sumida T.; Yanagawa H.; Doi N. In Vitro Selection of Fab Fragments by MRNA Display and Gene-Linking Emulsion PCR. J. Nucleic Acids 2012, 2012, 371379. 10.1155/2012/371379. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sumida T.; Doi N.; Yanagawa H. Bicistronic DNA Display for in Vitro Selection of Fab Fragments. Nucleic Acids Res. 2009, 37 (22), e147 10.1093/nar/gkp776. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ueno S.; Yoshida S.; Mondal A.; Nishina K.; Koyama M.; Sakata I.; Miura K.; Hayashi Y.; Nemoto N.; Nishigaki K.; Sakai T. In Vitro Selection of a Peptide Antagonist of Growth Hormone Secretagogue Receptor Using CDNA Display. Proc. Natl. Acad. Sci. U. S. A. 2012, 109 (28), 11121–11126. 10.1073/pnas.1203561109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jalali-Yazdi F.; Huong Lai L.; Takahashi T. T.; Roberts R. W. High-throughput Measurement of Binding Kinetics by MRNA Display and Next-generation Sequencing. Angew. Chem., Int. Ed. Engl. 2016, 55 (12), 4007–4010. 10.1002/anie.201600077. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Seelig B.; Szostak J. W. Selection and Evolution of Enzymes from a Partially Randomized Non-Catalytic Scaffold. Nature 2007, 448 (7155), 828–831. 10.1038/nature06032. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tsuboyama K.; Dauparas J.; Chen J.; Laine E.; Mohseni Behbahani Y.; Weinstein J. J.; Mangan N. M.; Ovchinnikov S.; Rocklin G. J. Mega-Scale Experimental Analysis of Protein Folding Stability in Biology and Design. Nature 2023, 620 (7973), 434–444. 10.1038/s41586-023-06328-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Keefe A. D.; Szostak J. W. Functional Proteins from a Random-Sequence Library. Nature 2001, 410 (6829), 715–718. 10.1038/35070613. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yamaguchi J.; Naimuddin M.; Biyani M.; Sasaki T.; Machida M.; Kubo T.; Funatsu T.; Husimi Y.; Nemoto N. CDNA Display: A Novel Screening Method for Functional Disulfide-Rich Peptides by Solid-Phase Synthesis and Stabilization of MRNA-Protein Fusions. Nucleic Acids Res. 2009, 37 (16), e108 10.1093/nar/gkp514. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mochizuki Y.; Suzuki T.; Fujimoto K.; Nemoto N. A Versatile Puromycin-Linker Using CnvK for High-Throughput in Vitro Selection by CDNA Display. J. Biotechnol. 2015, 212, 174–180. 10.1016/j.jbiotec.2015.08.020. [DOI] [PubMed] [Google Scholar]
- Ishizawa T.; Kawakami T.; Reid P. C.; Murakami H. TRAP Display: A High-Speed Selection Method for the Generation of Functional Polypeptides. J. Am. Chem. Soc. 2013, 135 (14), 5433–5440. 10.1021/ja312579u. [DOI] [PubMed] [Google Scholar]
- Kondo T.; Eguchi M.; Kito S.; Fujino T.; Hayashi G.; Murakami H. CDNA TRAP Display for Rapid and Stable in Vitro Selection of Antibody-like Proteins. Chem. Commun. (Camb.) 2021, 57 (19), 2416–2419. 10.1039/D0CC07541H. [DOI] [PubMed] [Google Scholar]
- Zeng Y.; Woolley M.; Chockalingam K.; Thomas B.; Arora S.; Hook M.; Chen Z. Click Display: A Rapid and Efficient in Vitro Protein Display Method for Directed Evolution. Nucleic Acids Res. 2023, 51 (16), e89 10.1093/nar/gkad643. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Odegrip R.; Coomber D.; Eldridge B.; Hederer R.; Kuhlman P. A.; Ullman C.; FitzGerald K.; McGregor D. CIS Display: In Vitro Selection of Peptides from Libraries of Protein-DNA Complexes. Proc. Natl. Acad. Sci. U. S. A. 2004, 101 (9), 2806–2810. 10.1073/pnas.0400219101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Reiersen H.; Løbersli I.; Løset G. Å.; Hvattum E.; Simonsen B.; Stacy J. E.; McGregor D.; Fitzgerald K.; Welschof M.; Brekke O. H.; Marvik O. J. Covalent Antibody Display-an in Vitro Antibody-DNA Library Selection System. Nucleic Acids Res. 2005, 33 (1), e10 10.1093/nar/gni010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Doi N.; Yanagawa H. STABLE: Protein-DNA Fusion System for Screening of Combinatorial Protein Libraries in Vitro. FEBS Lett. 1999, 457 (2), 227–230. 10.1016/S0014-5793(99)01041-8. [DOI] [PubMed] [Google Scholar]
- Bertschinger J.; Neri D. Covalent DNA Display as a Novel Tool for Directed Evolution of Proteins in Vitro. Protein Eng. Des. Sel. 2004, 17 (9), 699–707. 10.1093/protein/gzh082. [DOI] [PubMed] [Google Scholar]
- Stein V.; Sielaff I.; Johnsson K.; Hollfelder F. A Covalent Chemical Genotype-Phenotype Linkage for in Vitro Protein Evolution. Chembiochem 2007, 8 (18), 2191–2194. 10.1002/cbic.200700459. [DOI] [PubMed] [Google Scholar]
- Houlihan G.; Gatti-Lafranconi P.; Lowe D.; Hollfelder F. Directed Evolution of Anti-HER2 DARPins by SNAP Display Reveals Stability/Function Trade-Offs in the Selection Process. Protein Eng. Des. Sel. 2015, 28 (9), 269–279. 10.1093/protein/gzv029. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gan R.; Yamanaka Y.; Kojima T.; Nakano H. Microbeads Display of Proteins Using Emulsion PCR and Cell-Free Protein Synthesis. Biotechnol. Prog. 2008, 24 (5), 1107–1114. 10.1002/btpr.43. [DOI] [PubMed] [Google Scholar]
- Nord O.; Uhlén M.; Nygren P.-A. Microbead Display of Proteins by Cell-Free Expression of Anchored DNA. J. Biotechnol. 2003, 106 (1), 1–13. 10.1016/j.jbiotec.2003.09.002. [DOI] [PubMed] [Google Scholar]
- Sepp A.; Tawfik D. S.; Griffiths A. D. Microbead Display by in Vitro Compartmentalisation: Selection for Binding Using Flow Cytometry. FEBS Lett. 2002, 532 (3), 455–458. 10.1016/S0014-5793(02)03740-7. [DOI] [PubMed] [Google Scholar]
- Fujii S.; Matsuura T.; Sunami T.; Nishikawa T.; Kazuta Y.; Yomo T. Liposome Display for in Vitro Selection and Evolution of Membrane Proteins. Nat. Protoc. 2014, 9 (7), 1578–1591. 10.1038/nprot.2014.107. [DOI] [PubMed] [Google Scholar]
- Gui Q.; Lawson T.; Shan S.; Yan L.; Liu Y. The Application of Whole Cell-Based Biosensors for Use in Environmental Analysis and in Medical Diagnostics. Sensors (Basel) 2017, 17 (7), 1623. 10.3390/s17071623. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Roy R.; Ray S.; Chowdhury A.; Anand R. Tunable Multiplexed Whole-Cell Biosensors as Environmental Diagnostics for Ppb-Level Detection of Aromatic Pollutants. ACS Sens. 2021, 6 (5), 1933–1939. 10.1021/acssensors.1c00329. [DOI] [PubMed] [Google Scholar]
- Saini R.; Hegde K.; Brar S. K.; Verma M.. Advances in Whole Cell-Based Biosensors in Environmental Monitoring. In Tools, Techniques and Protocols for Monitoring Environmental Contaminants; Elsevier, 2019; pp 263–284. 10.1016/b978-0-12-814679-8.00013-3. [DOI] [Google Scholar]
- Kylilis N.; Riangrungroj P.; Lai H.-E.; Salema V.; Fernández L. Á.; Stan G.-B. V.; Freemont P. S.; Polizzi K. M. Whole-Cell Biosensor with Tunable Limit of Detection Enables Low-Cost Agglutination Assays for Medical Diagnostic Applications. ACS Sens. 2019, 4 (2), 370–378. 10.1021/acssensors.8b01163. [DOI] [PubMed] [Google Scholar]
- Wang T.; Liang C.; Xing W.; Wu W.; Hou Y.; Zhang L.; Xiao S.; Xu H.; An Y.; Zheng M.; Liu L.; Nie L. Transcriptional Factor Engineering in Microbes for Industrial Biotechnology. J. Chem. Technol. Biotechnol. 2020, 95 (12), 3071–3078. 10.1002/jctb.6512. [DOI] [Google Scholar]
- Thavarajah W.; Verosloff M. S.; Jung J. K.; Alam K. K.; Miller J. D.; Jewett M. C.; Young S. L.; Lucks J. B. A Primer on Emerging Field-Deployable Synthetic Biology Tools for Global Water Quality Monitoring. npj Clean Water 2020, 3 (1), 1–10. 10.1038/s41545-020-0064-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wan X.; Ho T. Y. H.; Wang B.. Engineering Prokaryote Synthetic Biology Biosensors. In Handbook of Cell Biosensors; Thouand G., Ed.; Springer International Publishing: Cham, 2019; pp 1–37. 10.1007/978-3-319-47405-2_131-1. [DOI] [Google Scholar]
- Zhang L.; Guo W.; Lu Y. Advances in Cell-Free Biosensors: Principle, Mechanism, and Applications. Biotechnol. J. 2020, 15 (9), e2000187 10.1002/biot.202000187. [DOI] [PubMed] [Google Scholar]
- Gräwe A.; Dreyer A.; Vornholt T.; Barteczko U.; Buchholz L.; Drews G.; Ho U. L.; Jackowski M. E.; Kracht M.; Lüders J.; Bleckwehl T.; Rositzka L.; Ruwe M.; Wittchen M.; Lutter P.; Müller K.; Kalinowski J. A Paper-Based, Cell-Free Biosensor System for the Detection of Heavy Metals and Date Rape Drugs. PLoS One 2019, 14 (3), e0210940 10.1371/journal.pone.0210940. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gupta S.; Sarkar S.; Katranidis A.; Bhattacharya J. Development of a Cell-Free Optical Biosensor for Detection of a Broad Range of Mercury Contaminants in Water: A Plasmid DNA-Based Approach. ACS Omega 2019, 4 (5), 9480–9487. 10.1021/acsomega.9b00205. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kawaguchi T.; Chen Y. P.; Norman R. S.; Decho A. W. Rapid Screening of Quorum-Sensing Signal N-Acyl Homoserine Lactones by an in Vitro Cell-Free Assay. Appl. Environ. Microbiol. 2008, 74 (12), 3667–3671. 10.1128/AEM.02869-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thavarajah W.; Silverman A. D.; Verosloff M. S.; Kelley-Loughnane N.; Jewett M. C.; Lucks J. B. Point-of-Use Detection of Environmental Fluoride via a Cell-Free Riboswitch-Based Biosensor. ACS Synth. Biol. 2020, 9, 10. 10.1021/acssynbio.9b00347. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiang Y. S.; Riedel T. E.; Popoola J. A.; Morrow B. R.; Cai S.; Ellington A. D.; Bhadra S. Portable Platform for Rapid In-Field Identification of Human Fecal Pollution in Water. Water Res. 2018, 131, 186–195. 10.1016/j.watres.2017.12.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Verosloff M.; Chappell J.; Perry K. L.; Thompson J. R.; Lucks J. B. PLANTDx: A Molecular Diagnostic for Point-of-Use Detection. ACS Synth. Biol. 2019, 8, 902. 10.1021/acssynbio.8b00526. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Beabout K.; Ehrenworth Breedon A. M.; Blum S. M.; Miklos A. E.; Lux M. W.; Chávez J. L.; Goodson M. S. Detection of Bile Acids in Complex Matrices Using a Transcription Factor-Based Biosensor. ACS Biomater. Sci. Eng. 2023, 9 (9), 5151–5162. 10.1021/acsbiomaterials.2c01006. [DOI] [PubMed] [Google Scholar]
- Soltani M.; Hunt J. P.; Bundy B. C. Rapid RNase Inhibitor Production to Enable Low-Cost, on-Demand Cell-Free Protein Synthesis Biosensor Use in Human Body Fluids. Biotechnol. Bioeng. 2021, 118 (10), 3973–3983. 10.1002/bit.27874. [DOI] [PubMed] [Google Scholar]
- Hunt J. P.; Zhao E. L.; Free T. J.; Soltani M.; Warr C. A.; Benedict A. B.; Takahashi M. K.; Griffitts J. S.; Pitt W. G.; Bundy B. C. Towards Detection of SARS-CoV-2 RNA in Human Saliva: A Paper-Based Cell-Free Toehold Switch Biosensor with a Visual Bioluminescent Output. N. Biotechnol. 2022, 66, 53–60. 10.1016/j.nbt.2021.09.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Voyvodic P. L.; Pandi A.; Koch M.; Conejero I.; Valjent E.; Courtet P.; Renard E.; Faulon J.-L.; Bonnet J. Plug-and-Play Metabolic Transducers Expand the Chemical Detection Space of Cell-Free Biosensors. Nat. Commun. 2019, 10 (1), 1697. 10.1038/s41467-019-09722-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Beabout K.; Bernhards C. B.; Thakur M.; Turner K. B.; Cole S. D.; Walper S. A.; Chávez J. L.; Lux M. W. Optimization of Heavy Metal Sensors Based on Transcription Factors and Cell-Free Expression Systems. ACS Synth. Biol. 2021, 10 (11), 3040–3054. 10.1021/acssynbio.1c00331. [DOI] [PubMed] [Google Scholar]
- Myhrvold C.; Freije C. A.; Gootenberg J. S.; Abudayyeh O. O.; Metsky H. C.; Durbin A. F.; Kellner M. J.; Tan A. L.; Paul L. M.; Parham L. A.; Garcia K. F.; Barnes K. G.; Chak B.; Mondini A.; Nogueira M. L.; Isern S.; Michael S. F.; Lorenzana I.; Yozwiak N. L.; MacInnis B. L.; Bosch I.; Gehrke L.; Zhang F.; Sabeti P. C. Field-Deployable Viral Diagnostics Using CRISPR-Cas13. Science 2018, 360 (6387), 444–448. 10.1126/science.aas8836. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gootenberg J. S.; Abudayyeh O. O.; Lee J. W.; Essletzbichler P.; Dy A. J.; Joung J.; Verdine V.; Donghia N.; Daringer N. M.; Freije C. A.; Myhrvold C.; Bhattacharyya R. P.; Livny J.; Regev A.; Koonin E. V.; Hung D. T.; Sabeti P. C.; Collins J. J.; Zhang F. Nucleic Acid Detection with CRISPR-Cas13a/C2c2. Science 2017, 356 (6336), 438–442. 10.1126/science.aam9321. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Chen J. S.; Ma E.; Harrington L. B.; Da Costa M.; Tian X.; Palefsky J. M.; Doudna J. A. CRISPR-Cas12a Target Binding Unleashes Indiscriminate Single-Stranded DNase Activity. Science 2018, 360 (6387), 436–439. 10.1126/science.aar6245. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lim H. J.; Jang Y. J.; Lee K.-H.; Kim D.-M. Translational Detection of Nonproteinogenic Amino Acids Using an Engineered Complementary Cell-Free Protein Synthesis Assay. Anal. Chem. 2020, 92 (17), 11505–11510. 10.1021/acs.analchem.0c01978. [DOI] [PubMed] [Google Scholar]
- Jang Y.-J.; Lee K.-H.; Yoo T. H.; Kim D.-M. Complementary Cell-Free Translational Assay for Quantification of Amino Acids. Anal. Chem. 2017, 89 (18), 9638–9642. 10.1021/acs.analchem.7b01956. [DOI] [PubMed] [Google Scholar]
- Catherine C.; Oh S. J.; Lee K.-H.; Min S.-E.; Won J.-I.; Yun H.; Kim D.-M. Engineering Thermal Properties of Elastin-like Polypeptides by Incorporation of Unnatural Amino Acids in a Cell-Free Protein Synthesis System. Biotechnol. Bioprocess Eng. 2015, 20 (3), 417–422. 10.1007/s12257-015-0190-1. [DOI] [Google Scholar]
- Lee Y. J.; Lee S.; Kim D.-M.. Translational Detection of Indole by Complementary Cell-Free Protein Synthesis Assay. Front. Bioeng. Biotechnol. 2022, 10. 10.3389/fbioe.2022.900162. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cho I.-H.; Kim D. H.; Park S.. Electrochemical Biosensors: Perspective on Functional Nanomaterials for on-Site Analysis. Biomater. Res. 2020, 24. 10.1186/s40824-019-0181-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Patino Diaz A.; Bracaglia S.; Ranallo S.; Patino T.; Porchetta A.; Ricci F. Programmable Cell-Free Transcriptional Switches for Antibody Detection. J. Am. Chem. Soc. 2022, 144 (13), 5820–5826. 10.1021/jacs.1c11706. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu Y.; Zeng H.; Wu Q.; Jiang X.; Duan C.; Long J.; Chen M.; Yang X. A Sensing Strategy Combining T7 Promoter-Contained DNA Probe with CRISPR/Cas13a for Detection of Bacteria and Human Methyltransferase. Anal. Chim. Acta 2022, 1227, 340266. 10.1016/j.aca.2022.340266. [DOI] [PubMed] [Google Scholar]
- Lentini R.; Forlin M.; Martini L.; Del Bianco C.; Spencer A. C.; Torino D.; Mansy S. S. Fluorescent Proteins and in Vitro Genetic Organization for Cell-Free Synthetic Biology. ACS Synth. Biol. 2013, 2 (9), 482–489. 10.1021/sb400003y. [DOI] [PubMed] [Google Scholar]
- Mathur D.; Thakur M.; Díaz S. A.; Susumu K.; Stewart M. H.; Oh E.; Walper S. A.; Medintz I. L. Hybrid Nucleic Acid-Quantum Dot Assemblies as Multiplexed Reporter Platforms for Cell-Free Transcription Translation-Based Biosensors. ACS Synth. Biol. 2022, 11 (12), 4089–4102. 10.1021/acssynbio.2c00394. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Morey K.; Thomas-Fenderson T.; Watson A.; Sebesta J.; Peebles C.; Gentry-Weeks C.. Toehold Switch plus Signal Amplification Enables Rapid Detection. Biotechnol. J. 2023, 18. 10.1002/biot.202200607. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Arce A.; Guzman Chavez F.; Gandini C.; Puig J.; Matute T.; Haseloff J.; Dalchau N.; Molloy J.; Pardee K.; Federici F.. Decentralizing Cell-Free RNA Sensing with the Use of Low-Cost Cell Extracts. Front. Bioeng. Biotechnol. 2021, 9. 10.3389/fbioe.2021.727584. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sharpes C. E.; McManus J. B.; Blum S. M.; Mgboji G. E.; Lux M. W. Assessment of Colorimetric Reporter Enzymes in the PURE System. ACS Synth. Biol. 2021, 10 (11), 3205–3208. 10.1021/acssynbio.1c00360. [DOI] [PubMed] [Google Scholar]
- Sankar K.; Baer R.; Grazon C.; Sabatelle R. C.; Lecommandoux S.; Klapperich C. M.; Galagan J. E.; Grinstaff M. W. An Allosteric Transcription Factor DNA-Binding Electrochemical Biosensor for Progesterone. ACS Sens. 2022, 7 (4), 1132–1137. 10.1021/acssensors.2c00133. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Sadat Mousavi P.; Smith S. J.; Chen J. B.; Karlikow M.; Tinafar A.; Robinson C.; Liu W.; Ma D.; Green A. A.; Kelley S. O.; Pardee K. A Multiplexed, Electrochemical Interface for Gene-Circuit-Based Sensors. Nat. Chem. 2020, 12 (1), 48–55. 10.1038/s41557-019-0366-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yu P.; Lei C.; Nie Z. Integration of Electrochemical Interface and Cell-Free Synthetic Biology for Biosensing. J. Electroanal. Chem. (Lausanne Switz) 2022, 911, 116209. 10.1016/j.jelechem.2022.116209. [DOI] [Google Scholar]
- Jang Y.-J.; Lee K.-H.; Yoo T. H.; Kim D.-M. Interfacing a Personal Glucose Meter with Cell-Free Protein Synthesis for Rapid Analysis of Amino Acids. Anal. Chem. 2019, 91 (3), 2531–2535. 10.1021/acs.analchem.8b05526. [DOI] [PubMed] [Google Scholar]
- Amalfitano E.; Karlikow M.; Norouzi M.; Jaenes K.; Cicek S.; Masum F.; Sadat Mousavi P.; Guo Y.; Tang L.; Sydor A.; Ma D.; Pearson J. D.; Trcka D.; Pinette M.; Ambagala A.; Babiuk S.; Pickering B.; Wrana J.; Bremner R.; Mazzulli T.; Sinton D.; Brumell J. H.; Green A. A.; Pardee K.. A Glucose Meter Interface for Point-of-Care Gene Circuit-Based Diagnostics. Nat. Commun. 2021, 12. 10.1038/s41467-020-20639-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Du Y.; Hughes R. A.; Bhadra S.; Jiang Y. S.; Ellington A. D.; Li B. A Sweet Spot for Molecular Diagnostics: Coupling Isothermal Amplification and Strand Exchange Circuits to Glucometers. Sci. Rep. 2015, 5 (1), 11039. 10.1038/srep11039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Silverman A. D.; Akova U.; Alam K. K.; Jewett M. C.; Lucks J. B. Design and Optimization of a Cell-Free Atrazine Biosensor. ACS Synth. Biol. 2020, 9 (3), 671–677. 10.1021/acssynbio.9b00388. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang X.; Zhu K.; Chen D.; Wang J.; Wang X.; Xu A.; Wu L.; Li L.; Chen S. Monitoring Arsenic Using Genetically Encoded Biosensors in Vitro: The Role of Evolved Regulatory Genes. Ecotoxicol. Environ. Saf. 2021, 207, 111273. 10.1016/j.ecoenv.2020.111273. [DOI] [PubMed] [Google Scholar]
- Zhang Y.; Zhao C.; Bi H.; Zhang X.; Xue B.; Li C.; Wang S.; Yang X.; Qiu Z.; Wang J.; Shen Z. A Cell-Free Paper-Based Biosensor Dependent on Allosteric Transcription Factors (ATFs) for on-Site Detection of Harmful Metals Hg2+ and Pb2+ in Water. J. Hazard. Mater. 2022, 438, 129499. 10.1016/j.jhazmat.2022.129499. [DOI] [PubMed] [Google Scholar]
- Pellinen T.; Huovinen T.; Karp M. A Cell-Free Biosensor for the Detection of Transcriptional Inducers Using Firefly Luciferase as a Reporter. Anal. Biochem. 2004, 330 (1), 52–57. 10.1016/j.ab.2004.03.064. [DOI] [PubMed] [Google Scholar]
- Lopreside A.; Wan X.; Michelini E.; Roda A.; Wang B. Comprehensive Profiling of Diverse Genetic Reporters with Application to Whole-Cell and Cell-Free Biosensors. Anal. Chem. 2019, 91 (23), 15284–15292. 10.1021/acs.analchem.9b04444. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhang P.; Feng H.; Yang J.; Jiang H.; Zhou H.; Lu Y. Detection of Inorganic Ions and Organic Molecules with Cell-Free Biosensing Systems. J. Biotechnol. 2019, 300, 78–86. 10.1016/j.jbiotec.2019.05.011. [DOI] [PubMed] [Google Scholar]
- Jung J. K.; Rasor B. J.; Rybnicky G. A.; Silverman A. D.; Standeven J.; Kuhn R.; Granito T.; Ekas H. M.; Wang B. M.; Karim A. S.; Lucks J. B.; Jewett M. C. At-Home, Cell-Free Synthetic Biology Education Modules for Transcriptional Regulation and Environmental Water Quality Monitoring. ACS Synth. Biol. 2023, 12 (10), 2909–2921. 10.1021/acssynbio.3c00223. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jung J. K.; Archuleta C. M.; Alam K. K.; Lucks J. B. Programming Cell-Free Biosensors with DNA Strand Displacement Circuits. Nat. Chem. Biol. 2022, 18 (4), 385–393. 10.1038/s41589-021-00962-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Karig D. K.; Iyer S.; Simpson M. L.; Doktycz M. J. Expression Optimization and Synthetic Gene Networks in Cell-Free Systems. Nucleic Acids Res. 2012, 40 (8), 3763–3774. 10.1093/nar/gkr1191. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yang Y.-H.; Kim T.-W.; Park S.-H.; Lee K.; Park H.-Y.; Song E.; Joo H.-S.; Kim Y.-G.; Hahn J.-S.; Kim B.-G. Cell-Free Escherichia Coli-Based System to Screen for Quorum-Sensing Molecules Interacting with Quorum Receptor Proteins of Streptomyces Coelicolor. Appl. Environ. Microbiol. 2009, 75 (19), 6367–6372. 10.1128/AEM.00019-09. [DOI] [PMC free article] [PubMed] [Google Scholar]
- McKeague M.; McConnell E. M.; Cruz-Toledo J.; Bernard E. D.; Pach A.; Mastronardi E.; Zhang X.; Beking M.; Francis T.; Giamberardino A.; Cabecinha A.; Ruscito A.; Aranda-Rodriguez R.; Dumontier M.; DeRosa M. C. Analysis of in Vitro Aptamer Selection Parameters. J. Mol. Evol. 2015, 81 (5–6), 150–161. 10.1007/s00239-015-9708-6. [DOI] [PubMed] [Google Scholar]
- Barrick J. E.; Breaker R. R. The Distributions, Mechanisms, and Structures of Metabolite-Binding Riboswitches. Genome Biol. 2007, 8 (11), R239. 10.1186/gb-2007-8-11-r239. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Garst A. D.; Edwards A. L.; Batey R. T. Riboswitches: Structures and Mechanisms. Cold Spring Harb. Perspect. Biol. 2011, 3 (6), a003533–a003533. 10.1101/cshperspect.a003533. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cho E. J.; Lee J.-W.; Ellington A. D. Applications of Aptamers as Sensors. Annu. Rev. Anal. Chem. (Palo Alto Calif.) 2009, 2 (1), 241–264. 10.1146/annurev.anchem.1.031207.112851. [DOI] [PubMed] [Google Scholar]
- Tabuchi T.; Yokobayashi Y. Cell-Free Riboswitches. RSC Chem. Biol. 2021, 2 (5), 1430–1440. 10.1039/D1CB00138H. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Adamala K. P.; Martin-Alarcon D. A.; Guthrie-Honea K. R.; Boyden E. S. Engineering Genetic Circuit Interactions within and between Synthetic Minimal Cells. Nat. Chem. 2017, 9 (5), 431–439. 10.1038/nchem.2644. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martini L.; Mansy S. S.. Measuring Riboswitch Activity in Vitro and in Artificial Cells with Purified Transcription-Translation Machinery. In Methods Mol. Biol.; Methods in Molecular Biology (Clifton, N.J.); Humana Press: Totowa, NJ, 2014; pp 153–164. 10.1007/978-1-62703-755-6_11. [DOI] [PubMed] [Google Scholar]
- Martini L.; Mansy S. S. Cell-like Systems with Riboswitch Controlled Gene Expression. Chem. Commun. (Camb.) 2011, 47 (38), 10734. 10.1039/c1cc13930d. [DOI] [PubMed] [Google Scholar]
- Iyer S.; Doktycz M. J. Thrombin-Mediated Transcriptional Regulation Using DNA Aptamers in DNA-Based Cell-Free Protein Synthesis. ACS Synth. Biol. 2014, 3 (6), 340–346. 10.1021/sb4000756. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang J.; Yang L.; Cui X.; Zhang Z.; Dong L.; Guan N. A DNA Bubble-Mediated Gene Regulation System Based on Thrombin-Bound DNA Aptamers. ACS Synth. Biol. 2017, 6 (5), 758–765. 10.1021/acssynbio.6b00391. [DOI] [PubMed] [Google Scholar]
- Boyd M. A.; Thavarajah W.; Lucks J. B.; Kamat N. P.. Robust and Tunable Performance of a Cell-Free Biosensor Encapsulated in Lipid Vesicles. Sci. Adv. 2023, 9 ( (1), ). 10.1126/sciadv.add6605. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Vezeau G. E.; Gadila L. R.; Salis H. M.. Automated Design of Protein-Binding Riboswitches for Sensing Human Biomarkers in a Cell-Free Expression System. Nat. Commun. 2023, 14 ( (1), ). 10.1038/s41467-023-38098-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dong X.; Cheng Q.; Qi S.; Qin M.; Ding N.; Sun Y.; Xia Y.; Zhang Y.; Wang Z. Broad-Spectrum Detection of Tetracyclines by Riboswitch-Based Cell-Free Expression Biosensing. J. Agric. Food Chem. 2023, 71 (25), 9886–9895. 10.1021/acs.jafc.3c02137. [DOI] [PubMed] [Google Scholar]
- Tabuchi T.; Yokobayashi Y. High-Throughput Screening of Cell-Free Riboswitches by Fluorescence-Activated Droplet Sorting. Nucleic Acids Res. 2022, 50 (6), 3535–3550. 10.1093/nar/gkac152. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ellington A. D.; Szostak J. W. In Vitro Selection of RNA Molecules That Bind Specific Ligands. Nature 1990, 346 (6287), 818–822. 10.1038/346818a0. [DOI] [PubMed] [Google Scholar]
- Dunn M. R.; Jimenez R. M.; Chaput J. C.. Analysis of Aptamer Discovery and Technology. Nat. Rev. Chem. 2017, 1 ( (10), ). 10.1038/s41570-017-0076. [DOI] [Google Scholar]
- Findeiß S.; Etzel M.; Will S.; Mörl M.; Stadler P. Design of Artificial Riboswitches as Biosensors. Sensors (Basel) 2017, 17 (9), 1990. 10.3390/s17091990. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Thavarajah W.; Owuor P. M.; Awuor D. R.; Kiprotich K.; Aggarwal R.; Lucks J. B.; Young S. L.. The Accuracy and Usability of Point-of-Use Fluoride Biosensors in Rural Kenya. Npj Clean Water 2023, 6 ( (1), ). 10.1038/s41545-023-00221-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- de Puig H.; Bosch I.; Collins J. J.; Gehrke L. Point-of-Care Devices to Detect Zika and Other Emerging Viruses. Annu. Rev. Biomed. Eng. 2020, 22 (1), 371–386. 10.1146/annurev-bioeng-060418-052240. [DOI] [PubMed] [Google Scholar]
- Köksaldı I. Ç.; Köse S.; Ahan R. E.; Hacıosmanoǧlu N.; Şahin Kehribar E.; Güngen M. A.; Baştuǧ A.; Dinç B.; Bodur H.; Özkul A.; Şeker U. Ö. Ş. SARS-CoV-2 Detection with DE Novo-Designed Synthetic Riboregulators. Anal. Chem. 2021, 93 (28), 9719–9727. 10.1021/acs.analchem.1c00886. [DOI] [PubMed] [Google Scholar]
- Carr A. R.; Dopp J. L.; Wu K.; Sadat Mousavi P.; Jo Y. R.; McNeley C. E.; Lynch Z. T.; Pardee K.; Green A. A.; Reuel N. F. Toward Mail-in-Sensors for SARS-CoV-2 Detection: Interfacing Gel Switch Resonators with Cell-Free Toehold Switches. ACS Sens. 2022, 7 (3), 806–815. 10.1021/acssensors.1c02450. [DOI] [PubMed] [Google Scholar]
- Cao M.; Sun Q.; Zhang X.; Ma Y.; Wang J. Detection and Differentiation of Respiratory Syncytial Virus Subgroups A and B with Colorimetric Toehold Switch Sensors in a Paper-Based Cell-Free System. Biosens. Bioelectron. 2021, 182, 113173. 10.1016/j.bios.2021.113173. [DOI] [PubMed] [Google Scholar]
- Ma D.; Shen L.; Wu K.; Diehnelt C. W.; Green A. A. Low-Cost Detection of Norovirus Using Paper-Based Cell-Free Systems and Synbody-Based Viral Enrichment. Synth. Biol. 2018, 3 (1), ysy018. 10.1093/synbio/ysy018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kang X.; Zhao C.; Chen S.; Zhang X.; Xue B.; Li C.; Wang S.; Yang X.; Xia Z.; Xu Y.; Huang Y.; Qiu Z.; Li C.; Wang J.; Pang J.; Shen Z. Development of a Cell-Free Toehold Switch for Hepatitis A Virus Type I on-Site Detection. Anal. Methods 2023, 15 (43), 5813–5822. 10.1039/D3AY01408H. [DOI] [PubMed] [Google Scholar]
- Takahashi M. K.; Tan X.; Dy A. J.; Braff D.; Akana R. T.; Furuta Y.; Donghia N.; Ananthakrishnan A.; Collins J. J. A Low-Cost Paper-Based Synthetic Biology Platform for Analyzing Gut Microbiota and Host Biomarkers. Nat. Commun. 2018, 9 (1), 3347. 10.1038/s41467-018-05864-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wang Y.; Liu J.; Zhou H. Visual Detection of Cucumber Green Mottle Mosaic Virus Based on Terminal Deoxynucleotidyl Transferase Coupled with DNAzymes Amplification. Sensors (Basel) 2019, 19 (6), 1298. 10.3390/s19061298. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Green A. A.; Silver P. A.; Collins J. J.; Yin P. Toehold Switches: De-Novo-Designed Regulators of Gene Expression. Cell 2014, 159 (4), 925–939. 10.1016/j.cell.2014.10.002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- To A. C.-Y.; Chu D. H.-T.; Wang A. R.; Li F. C.-Y.; Chiu A. W.-O.; Gao D. Y.; Choi C. H. J.; Kong S.-K.; Chan T.-F.; Chan K.-M.; Yip K. Y. A Comprehensive Web Tool for Toehold Switch Design. Bioinformatics 2018, 34 (16), 2862–2864. 10.1093/bioinformatics/bty216. [DOI] [PubMed] [Google Scholar]
- Guatelli J. C.; Whitfield K. M.; Kwoh D. Y.; Barringer K. J.; Richman D. D.; Gingeras T. R. Isothermal, in Vitro Amplification of Nucleic Acids by a Multienzyme Reaction Modeled after Retroviral Replication. Proc. Natl. Acad. Sci. U. S. A. 1990, 87 (5), 1874–1878. 10.1073/pnas.87.5.1874. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Notomi T.; Okayama H.; Masubuchi H.; Yonekawa T.; Watanabe K.; Amino N.; Hase T. Loop-Mediated Isothermal Amplification of DNA. Nucleic Acids Res. 2000, 28 (12), 63e 10.1093/nar/28.12.e63. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Piepenburg O.; Williams C. H.; Stemple D. L.; Armes N. A. DNA Detection Using Recombination Proteins. PLoS Biol. 2006, 4 (7), e204 10.1371/journal.pbio.0040204. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Boonbanjong P.; Treerattrakoon K.; Waiwinya W.; Pitikultham P.; Japrung D. Isothermal Amplification Technology for Disease Diagnosis. Biosensors (Basel) 2022, 12 (9), 677. 10.3390/bios12090677. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kinpara T.; Mizuno R.; Murakami Y.; Kobayashi M.; Yamaura S.; Hasan Q.; Morita Y.; Nakano H.; Yamane T.; Tamiya E. A Picoliter Chamber Array for Cell-Free Protein Synthesis. J. Biochem. 2004, 136 (2), 149–154. 10.1093/jb/mvh102. [DOI] [PubMed] [Google Scholar]
- Hansen M. M. K.; Paffenholz S.; Foschepoth D.; Heus H. A.; Thiele J.; Huck W. T. S. Cell-like Nanostructured Environments Alter Diffusion and Reaction Kinetics in Cell-Free Gene Expression. Chembiochem 2016, 17 (3), 228–232. 10.1002/cbic.201500560. [DOI] [PubMed] [Google Scholar]
- Marshall R.; Maxwell C. S.; Collins S. P.; Jacobsen T.; Luo M. L.; Begemann M. B.; Gray B. N.; January E.; Singer A.; He Y.; Beisel C. L.; Noireaux V. Rapid and Scalable Characterization of CRISPR Technologies Using an E. Coli Cell-Free Transcription-Translation System. Mol. Cell 2018, 69 (1), 146–157. e3 10.1016/j.molcel.2017.12.007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Borkowski O.; Bricio C.; Murgiano M.; Rothschild-Mancinelli B.; Stan G.-B.; Ellis T.. Cell-Free Prediction of Protein Expression Costs for Growing Cells. Nat. Commun. 2018, 9 ( (1), ). 10.1038/s41467-018-03970-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takahashi M. K.; Chappell J.; Hayes C. A.; Sun Z. Z.; Kim J.; Singhal V.; Spring K. J.; Al-Khabouri S.; Fall C. P.; Noireaux V.; Murray R. M.; Lucks J. B. Rapidly Characterizing the Fast Dynamics of RNA Genetic Circuitry with Cell-Free Transcription-Translation (TX-TL) Systems. ACS Synth. Biol. 2015, 4 (5), 503–515. 10.1021/sb400206c. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dittrich P. S.; Jahnz M.; Schwille P. A New Embedded Process for Compartmentalized Cell-Free Protein Expression and on-Line Detection in Microfluidic Devices. Chembiochem 2005, 6 (5), 811–814. 10.1002/cbic.200400321. [DOI] [PubMed] [Google Scholar]
- Gan R.; Cabezas M. D.; Pan M.; Zhang H.; Hu G.; Clark L. G.; Jewett M. C.; Nicol R. High-Throughput Regulatory Part Prototyping and Analysis by Cell-Free Protein Synthesis and Droplet Microfluidics. ACS Synth. Biol. 2022, 11, 2108. 10.1021/acssynbio.2c00050. [DOI] [PubMed] [Google Scholar]
- Kahn J. S.; Ruiz R. C. H.; Sureka S.; Peng S.; Derrien T. L.; An D.; Luo D. DNA Microgels as a Platform for Cell-Free Protein Expression and Display. Biomacromolecules 2016, 17 (6), 2019–2026. 10.1021/acs.biomac.6b00183. [DOI] [PubMed] [Google Scholar]
- Vezeau G. E.; Salis H. M. Tuning Cell-Free Composition Controls the Time Delay, Dynamics, and Productivity of TX-TL Expression. ACS Synth. Biol. 2021, 10 (10), 2508–2519. 10.1021/acssynbio.1c00136. [DOI] [PubMed] [Google Scholar]
- Brophy J. A. N.; Voigt C. A. Principles of Genetic Circuit Design. Nat. Methods 2014, 11 (5), 508–520. 10.1038/nmeth.2926. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lehr F.-X.; Hanst M.; Vogel M.; Kremer J.; Göringer H. U.; Suess B.; Koeppl H. Cell-Free Prototyping of AND-Logic Gates Based on Heterogeneous RNA Activators. ACS Synth. Biol. 2019, 8 (9), 2163–2173. 10.1021/acssynbio.9b00238. [DOI] [PubMed] [Google Scholar]
- Noireaux V.; Bar-Ziv R.; Libchaber A. Principles of Cell-Free Genetic Circuit Assembly. Proc. Natl. Acad. Sci. U. S. A. 2003, 100 (22), 12672–12677. 10.1073/pnas.2135496100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Takahashi M. K.; Hayes C. A.; Chappell J.; Sun Z. Z.; Murray R. M.; Noireaux V.; Lucks J. B. Characterizing and Prototyping Genetic Networks with Cell-Free Transcription-Translation Reactions. Methods 2015, 86, 60–72. 10.1016/j.ymeth.2015.05.020. [DOI] [PubMed] [Google Scholar]
- Patel A.; Murray R. M.; Sen S. Assessment of Robustness to Temperature in a Negative Feedback Loop and a Feedforward Loop. ACS Synth. Biol. 2020, 9 (7), 1581–1590. 10.1021/acssynbio.0c00023. [DOI] [PubMed] [Google Scholar]
- Pieters P. A.; Nathalia B. L.; van der Linden A. J.; Yin P.; Kim J.; Huck W. T. S.; de Greef T. F. A. Cell-Free Characterization of Coherent Feed-Forward Loop-Based Synthetic Genetic Circuits. ACS Synth. Biol. 2021, 10 (6), 1406–1416. 10.1021/acssynbio.1c00024. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Guo S.; Murray R. M. Construction of Incoherent Feedforward Loop Circuits in a Cell-Free System and in Cells. ACS Synth. Biol. 2019, 8 (3), 606–610. 10.1021/acssynbio.8b00493. [DOI] [PubMed] [Google Scholar]
- Westbrook A.; Tang X.; Marshall R.; Maxwell C. S.; Chappell J.; Agrawal D. K.; Dunlop M. J.; Noireaux V.; Beisel C. L.; Lucks J.; Franco E. Distinct Timescales of RNA Regulators Enable the Construction of a Genetic Pulse Generator. Biotechnol. Bioeng. 2019, 116 (5), 1139–1151. 10.1002/bit.26918. [DOI] [PubMed] [Google Scholar]
- Hu C. Y.; Takahashi M. K.; Zhang Y.; Lucks J. B. Engineering a Functional Small RNA Negative Autoregulation Network with Model-Guided Design. ACS Synth. Biol. 2018, 7 (6), 1507–1518. 10.1021/acssynbio.7b00440. [DOI] [PubMed] [Google Scholar]
- Niederholtmeyer H.; Stepanova V.; Maerkl S. J. Implementation of Cell-Free Biological Networks at Steady State. Proc. Natl. Acad. Sci. U. S. A. 2013, 110 (40), 15985–15990. 10.1073/pnas.1311166110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Niederholtmeyer H.; Sun Z. Z.; Hori Y.; Yeung E.; Verpoorte A.; Murray R. M.; Maerkl S. J. Rapid Cell-Free Forward Engineering of Novel Genetic Ring Oscillators. Elife 2015, 4, e09771 10.7554/eLife.09771. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Yelleswarapu M.; van der Linden A. J.; van Sluijs B.; Pieters P. A.; Dubuc E.; de Greef T. F. A.; Huck W. T. S. Sigma Factor-Mediated Tuning of Bacterial Cell-Free Synthetic Genetic Oscillators. ACS Synth. Biol. 2018, 7 (12), 2879–2887. 10.1021/acssynbio.8b00300. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rhea K. A.; McDonald N. D.; Cole S. D.; Noireaux V.; Lux M. W.; Buckley P. E.. Variability in Cell-Free Expression Reactions Can Impact Qualitative Genetic Circuit Characterization. Synth. Biol. 2022, 7 ( (1), ). 10.1093/synbio/ysac011. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bell E. L.; Finnigan W.; France S. P.; Green A. P.; Hayes M. A.; Hepworth L. J.; Lovelock S. L.; Niikura H.; Osuna S.; Romero E.; Ryan K. S.; Turner N. J.; Flitsch S. L. Biocatalysis. Nat. Rev. Methods Primers 2021, 1 (1), 1–21. 10.1038/s43586-021-00044-z. [DOI] [Google Scholar]
- France S. P.; Lewis R. D.; Martinez C. A. The Evolving Nature of Biocatalysis in Pharmaceutical Research and Development. JACS Au 2023, 3 (3), 715–735. 10.1021/jacsau.2c00712. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kwon Y.-C.; Song J.-K.; Kim D.-M.. Cloning-Independent Expression and Screening of Enzymes Using Cell-Free Protein Synthesis Systems. In Methods Mol. Biol.; Methods in Molecular Biology (Clifton, N.J.); Humana Press: Totowa, NJ, 2014; pp 97–108. 10.1007/978-1-62703-782-2_6. [DOI] [PubMed] [Google Scholar]
- Rolf J.; Ngo A. C. R.; Lütz S.; Tischler D.; Rosenthal K.. Cell-free Protein Synthesis for the Screening of Novel Azoreductases and Their Preferred Electron Donor. Chembiochem 2022, 23 ( (15), ). 10.1002/cbic.202200121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rolf J.; Rosenthal K.; Lütz S. Application of Cell-Free Protein Synthesis for Faster Biocatalyst Development. Catalysts 2019, 9 (2), 190. 10.3390/catal9020190. [DOI] [Google Scholar]
- Lim H. J.; Kim D.-M.. Cell-Free Metabolic Engineering: Recent Developments and Future Prospects. Methods Protoc. 2019, 2 ( (2), ). 33. 10.3390/mps2020033. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Karim A. S.; Jewett M. C. A Cell-Free Framework for Rapid Biosynthetic Pathway Prototyping and Enzyme Discovery. Metab. Eng. 2016, 36, 116–126. 10.1016/j.ymben.2016.03.002. [DOI] [PubMed] [Google Scholar]
- Bingham A. S.; Smith P. R.; Swartz J. R. Evolution of an [FeFe] Hydrogenase with Decreased Oxygen Sensitivity. Int. J. Hydrogen Energy 2012, 37 (3), 2965–2976. 10.1016/j.ijhydene.2011.02.048. [DOI] [Google Scholar]
- Ueno H.; Kato M.; Minagawa Y.; Hirose Y.; Noji H. Elucidation and Control of Low and High Active Populations of Alkaline Phosphatase Molecules for Quantitative Digital Bioassay. Protein Sci. 2021, 30 (8), 1628–1639. 10.1002/pro.4102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Mastrobattista E.; Taly V.; Chanudet E.; Treacy P.; Kelly B. T.; Griffiths A. D. High-Throughput Screening of Enzyme Libraries: In Vitro Evolution of a Beta-Galactosidase by Fluorescence-Activated Sorting of Double Emulsions. Chem. Biol. 2005, 12 (12), 1291–1300. 10.1016/j.chembiol.2005.09.016. [DOI] [PubMed] [Google Scholar]
- Woronoff G.; El Harrak A.; Mayot E.; Schicke O.; Miller O. J.; Soumillion P.; Griffiths A. D.; Ryckelynck M. New Generation of Amino Coumarin Methyl Sulfonate-Based Fluorogenic Substrates for Amidase Assays in Droplet-Based Microfluidic Applications. Anal. Chem. 2011, 83 (8), 2852–2857. 10.1021/ac200373n. [DOI] [PubMed] [Google Scholar]
- Zhu B.; Mizoguchi T.; Kojima T.; Nakano H. Ultra-High-Throughput Screening of an in Vitro-Synthesized Horseradish Peroxidase Displayed on Microbeads Using Cell Sorter. PLoS One 2015, 10 (5), e0127479 10.1371/journal.pone.0127479. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Körfer G.; Pitzler C.; Vojcic L.; Martinez R.; Schwaneberg U. In Vitro Flow Cytometry-Based Screening Platform for Cellulase Engineering. Sci. Rep. 2016, 6, 26128. 10.1038/srep26128. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fischlechner M.; Schaerli Y.; Mohamed M. F.; Patil S.; Abell C.; Hollfelder F. Evolution of Enzyme Catalysts Caged in Biomimetic Gel-Shell Beads. Nat. Chem. 2014, 6 (9), 791–796. 10.1038/nchem.1996. [DOI] [PubMed] [Google Scholar]
- Lindenburg L.; Hollfelder F. NAD-Display”: Ultrahigh-Throughput in Vitro Screening of NAD(H) Dehydrogenases Using Bead Display and Flow Cytometry. Angew. Chem., Int. Ed. Engl. 2021, 60 (16), 9015–9021. 10.1002/anie.202013486. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zinchenko A.; Devenish S. R. A.; Kintses B.; Colin P.-Y.; Fischlechner M.; Hollfelder F. One in a Million: Flow Cytometric Sorting of Single Cell-Lysate Assays in Monodisperse Picolitre Double Emulsion Droplets for Directed Evolution. Anal. Chem. 2014, 86 (5), 2526–2533. 10.1021/ac403585p. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Holland-Moritz D. A.; Wismer M. K.; Mann B. F.; Farasat I.; Devine P.; Guetschow E. D.; Mangion I.; Welch C. J.; Moore J. C.; Sun S.; Kennedy R. T. Mass Activated Droplet Sorting (MADS) Enables High-Throughput Screening of Enzymatic Reactions at Nanoliter Scale. Angew. Chem., Int. Ed. Engl. 2020, 59 (11), 4470–4477. 10.1002/anie.201913203. [DOI] [PubMed] [Google Scholar]
- Sunami T.; Matsuura T.; Suzuki H.; Yomo T.. Synthesis of Functional Proteins within Liposomes. In Methods Mol. Biol.; Methods in Molecular Biology (Clifton, N.J.); Humana Press: Totowa, NJ, 2010; pp 243–256. 10.1007/978-1-60327-331-2_20. [DOI] [PubMed] [Google Scholar]
- Schulz L.; Guo Z.; Zarzycki J.; Steinchen W.; Schuller J. M.; Heimerl T.; Prinz S.; Mueller-Cajar O.; Erb T. J.; Hochberg G. K. A. Evolution of Increased Complexity and Specificity at the Dawn of Form I Rubiscos. Science 2022, 378 (6616), 155–160. 10.1126/science.abq1416. [DOI] [PubMed] [Google Scholar]
- Lu H.; Diaz D. J.; Czarnecki N. J.; Zhu C.; Kim W.; Shroff R.; Acosta D. J.; Alexander B. R.; Cole H. O.; Zhang Y.; Lynd N. A.; Ellington A. D.; Alper H. S. Machine Learning-Aided Engineering of Hydrolases for PET Depolymerization. Nature 2022, 604 (7907), 662–667. 10.1038/s41586-022-04599-z. [DOI] [PubMed] [Google Scholar]
- Yeh A. H.-W.; Norn C.; Kipnis Y.; Tischer D.; Pellock S. J.; Evans D.; Ma P.; Lee G. R.; Zhang J. Z.; Anishchenko I.; Coventry B.; Cao L.; Dauparas J.; Halabiya S.; DeWitt M.; Carter L.; Houk K. N.; Baker D. De Novo Design of Luciferases Using Deep Learning. Nature 2023, 614 (7949), 774–780. 10.1038/s41586-023-05696-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ahsan A.; Wagner D.; Varaljay V. A.; Roman V.; Kelley-Loughnane N.; Reuel N. F.. Screening Putative Polyester Polyurethane Degrading Enzymes with Semi-Automated Cell-Free Expression and Nitrophenyl Probes. Synth. Biol. 2024, 9 ( (1), ). 10.1093/synbio/ysae005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bowie J. U.; Sherkhanov S.; Korman T. P.; Valliere M. A.; Opgenorth P. H.; Liu H. Synthetic Biochemistry: The Bio-Inspired Cell-Free Approach to Commodity Chemical Production. Trends Biotechnol. 2020, 38 (7), 766–778. 10.1016/j.tibtech.2019.12.024. [DOI] [PubMed] [Google Scholar]
- Dudley Q. M.; Karim A. S.; Jewett M. C. Cell-Free Metabolic Engineering: Biomanufacturing beyond the Cell. Biotechnol. J. 2015, 10 (1), 69–82. 10.1002/biot.201400330. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rollin J. A.; Bomble Y. J.; St John P. C.; Stark A. K. Biochemical Production with Purified Cell-Free Systems. Biochem. Eng. J. 2021, 166, 107002. 10.1016/j.bej.2018.07.016. [DOI] [Google Scholar]
- Guterl J.-K.; Garbe D.; Carsten J.; Steffler F.; Sommer B.; Reiße S.; Philipp A.; Haack M.; Rühmann B.; Koltermann A.; Kettling U.; Brück T.; Sieber V. Cell-Free Metabolic Engineering: Production of Chemicals by Minimized Reaction Cascades. ChemSusChem 2012, 5 (11), 2165–2172. 10.1002/cssc.201200365. [DOI] [PubMed] [Google Scholar]
- Sutiono S.; Pick A.; Sieber V. Converging Conversion - Using Promiscuous Biocatalysts for the Cell-Free Synthesis of Chemicals from Heterogeneous Biomass. Green Chem. 2021, 23 (10), 3656–3663. 10.1039/D0GC04288A. [DOI] [Google Scholar]
- Valliere M. A.; Korman T. P.; Arbing M. A.; Bowie J. U. A Bio-Inspired Cell-Free System for Cannabinoid Production from Inexpensive Inputs. Nat. Chem. Biol. 2020, 16 (12), 1427–1433. 10.1038/s41589-020-0631-9. [DOI] [PubMed] [Google Scholar]
- Sherkhanov S.; Korman T. P.; Chan S.; Faham S.; Liu H.; Sawaya M. R.; Hsu W.-T.; Vikram E.; Cheng T.; Bowie J. U. Isobutanol Production Freed from Biological Limits Using Synthetic Biochemistry. Nat. Commun. 2020, 11 (1), 4292. 10.1038/s41467-020-18124-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Korman T. P.; Opgenorth P. H.; Bowie J. U. A Synthetic Biochemistry Platform for Cell Free Production of Monoterpenes from Glucose. Nat. Commun. 2017, 8, 15526. 10.1038/ncomms15526. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Martin C. H.; Nielsen D. R.; Solomon K. V.; Prather K. L. J. Synthetic Metabolism: Engineering Biology at the Protein and Pathway Scales. Chem. Biol. 2009, 16 (3), 277–286. 10.1016/j.chembiol.2009.01.010. [DOI] [PubMed] [Google Scholar]
- Morgado G.; Gerngross D.; Roberts T. M.; Panke S. Synthetic Biology for Cell-Free Biosynthesis: Fundamentals of Designing Novel in Vitro Multi-Enzyme Reaction Networks. Adv. Biochem. Eng. Biotechnol. 2016, 162, 117–146. 10.1007/10_2016_13. [DOI] [PubMed] [Google Scholar]
- Erb T. J.; Jones P. R.; Bar-Even A. Synthetic Metabolism: Metabolic Engineering Meets Enzyme Design. Curr. Opin. Chem. Biol. 2017, 37, 56–62. 10.1016/j.cbpa.2016.12.023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hatzimanikatis V.; Li C.; Ionita J. A.; Henry C. S.; Jankowski M. D.; Broadbelt L. J. Exploring the Diversity of Complex Metabolic Networks. Bioinformatics 2005, 21 (8), 1603–1609. 10.1093/bioinformatics/bti213. [DOI] [PubMed] [Google Scholar]
- Ni Z.; Stine A. E.; Tyo K. E. J.; Broadbelt L. J. Curating a Comprehensive Set of Enzymatic Reaction Rules for Efficient Novel Biosynthetic Pathway Design. Metab. Eng. 2021, 65, 79–87. 10.1016/j.ymben.2021.02.006. [DOI] [PubMed] [Google Scholar]
- Bar-Even A.; Noor E.; Lewis N. E.; Milo R. Design and Analysis of Synthetic Carbon Fixation Pathways. Proc. Natl. Acad. Sci. U. S. A. 2010, 107 (19), 8889–8894. 10.1073/pnas.0907176107. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bogorad I. W.; Lin T.-S.; Liao J. C. Synthetic Non-Oxidative Glycolysis Enables Complete Carbon Conservation. Nature 2013, 502 (7473), 693–697. 10.1038/nature12575. [DOI] [PubMed] [Google Scholar]
- Clomburg J. M.; Vick J. E.; Blankschien M. D.; Rodríguez-Moyá M.; Gonzalez R. A Synthetic Biology Approach to Engineer a Functional Reversal of the β-Oxidation Cycle. ACS Synth. Biol. 2012, 1 (11), 541–554. 10.1021/sb3000782. [DOI] [PubMed] [Google Scholar]
- Tarasava K.; Lee S. H.; Chen J.; Köpke M.; Jewett M. C.; Gonzalez R. Reverse β-Oxidation Pathways for Efficient Chemical Production. J. Ind. Microbiol. Biotechnol. 2022, 10.1093/jimb/kuac003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Siegel J. B.; Smith A. L.; Poust S.; Wargacki A. J.; Bar-Even A.; Louw C.; Shen B. W.; Eiben C. B.; Tran H. M.; Noor E.; Gallaher J. L.; Bale J.; Yoshikuni Y.; Gelb M. H.; Keasling J. D.; Stoddard B. L.; Lidstrom M. E.; Baker D. Computational Protein Design Enables a Novel One-Carbon Assimilation Pathway. Proc. Natl. Acad. Sci. U. S. A. 2015, 112 (12), 3704–3709. 10.1073/pnas.1500545112. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schwander T.; Schada von Borzyskowski L.; Burgener S.; Cortina N. S.; Erb T. J. A Synthetic Pathway for the Fixation of Carbon Dioxide in Vitro. Science 2016, 354 (6314), 900–904. 10.1126/science.aah5237. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ninh P. H.; Honda K.; Sakai T.; Okano K.; Ohtake H. Assembly and Multiple Gene Expression of Thermophilic Enzymes in Escherichia Coli for in Vitro Metabolic Engineering. Biotechnol. Bioeng. 2015, 112 (1), 189–196. 10.1002/bit.25338. [DOI] [PubMed] [Google Scholar]
- Bat-Erdene U.; Billingsley J. M.; Turner W. C.; Lichman B. R.; Ippoliti F. M.; Garg N. K.; O’Connor S. E.; Tang Y. Cell-Free Total Biosynthesis of Plant Terpene Natural Products Using an Orthogonal Cofactor Regeneration System. ACS Catal. 2021, 11 (15), 9898–9903. 10.1021/acscatal.1c02267. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jo S.-M.; Wurm F. R.; Landfester K. Enzyme-Loaded Nanoreactors Enable the Continuous Regeneration of Nicotinamide Adenine Dinucleotide in Artificial Metabolisms. Angew. Chem., Int. Ed. Engl. 2021, 60 (14), 7728–7734. 10.1002/anie.202012023. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Black W. B.; Zhang L.; Mak W. S.; Maxel S.; Cui Y.; King E.; Fong B.; Sanchez Martinez A.; Siegel J. B.; Li H. Engineering a Nicotinamide Mononucleotide Redox Cofactor System for Biocatalysis. Nat. Chem. Biol. 2020, 16 (1), 87–94. 10.1038/s41589-019-0402-7. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Opgenorth P. H.; Korman T. P.; Bowie J. U. A Synthetic Biochemistry Module for Production of Bio-Based Chemicals from Glucose. Nat. Chem. Biol. 2016, 12 (6), 393–395. 10.1038/nchembio.2062. [DOI] [PubMed] [Google Scholar]
- Opgenorth P. H.; Korman T. P.; Bowie J. U. A Synthetic Biochemistry Molecular Purge Valve Module That Maintains Redox Balance. Nat. Commun. 2014, 5, 4113. 10.1038/ncomms5113. [DOI] [PubMed] [Google Scholar]
- Opgenorth P. H.; Korman T. P.; Iancu L.; Bowie J. U. A Molecular Rheostat Maintains ATP Levels to Drive a Synthetic Biochemistry System. Nat. Chem. Biol. 2017, 13 (9), 938–942. 10.1038/nchembio.2418. [DOI] [PubMed] [Google Scholar]
- Cui J.; Stevenson D.; Korosh T.; Amador-Noguez D.; Olson D. G.; Lynd L. R.. Developing a Cell-Free Extract Reaction (CFER) System in Clostridium Thermocellum to Identify Metabolic Limitations to Ethanol Production. Front. Energy Res. 2020, 8. 10.3389/fenrg.2020.00072. [DOI] [Google Scholar]
- Khattak W. A.; Ul-Islam M.; Ullah M. W.; Yu B.; Khan S.; Park J. K. Yeast Cell-Free Enzyme System for Bio-Ethanol Production at Elevated Temperatures. Process Biochem. 2014, 49 (3), 357–364. 10.1016/j.procbio.2013.12.019. [DOI] [Google Scholar]
- Schaffrath C.; Cobb S. L.; O’Hagan D. Cell-Free Biosynthesis of Fluoroacetate and 4-Fluorothreonine in Streptomyces Cattleya. Angew. Chem., Int. Ed. Engl. 2002, 41 (20), 3913–3915. . [DOI] [PubMed] [Google Scholar]
- Liu C.; Kelly G. T.; Watanabe C. M. H. In Vitro Biosynthesis of the Antitumor Agent Azinomycin B. Org. Lett. 2006, 8 (6), 1065–1068. 10.1021/ol052987l. [DOI] [PubMed] [Google Scholar]
- Jensen S. E.; Westlake D. W.; Bowers R. J.; Wolfe S. Cephalosporin Formation by Cell-Free Extracts from Streptomyces Clavuligerus. J. Antibiot. (Tokyo) 1982, 35 (10), 1351–1360. 10.7164/antibiotics.35.1351. [DOI] [PubMed] [Google Scholar]
- Jensen S. E.; Westlake D. W.; Bowers R. J.; Lyubechansky L.; Wolfe S. Synthesis of Benzylpenicillin by Cell-Free Extracts from Streptomyces Clavuligerus. J. Antibiot. (Tokyo) 1986, 39 (6), 822–826. 10.7164/antibiotics.39.822. [DOI] [PubMed] [Google Scholar]
- Malina H.; Tempete C.; Robert-Gero M. Biosynthesis of Sinefungin by Cell-Free Extract of Streptomyces Incarnatus NRRL 8089. J. Antibiot. (Tokyo) 1987, 40 (4), 505–511. 10.7164/antibiotics.40.505. [DOI] [PubMed] [Google Scholar]
- Kim J. K.; Kang H.-I.; Chae J. S.; Park Y.-H.; Choi Y.-J. Synthesis of Cefminox by Cell-Free Extracts OfStreptomyces Clavuligerus. FEMS Microbiol. Lett. 2000, 182 (2), 313–317. 10.1111/j.1574-6968.2000.tb08914.x. [DOI] [PubMed] [Google Scholar]
- Kominek L. A. Biosynthesis of Novobiocin by Streptomyces Niveus. Antimicrob. Agents Chemother. 1972, 1 (2), 123–134. 10.1128/AAC.1.2.123. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bujara M.; Schümperli M.; Billerbeck S.; Heinemann M.; Panke S. Exploiting Cell-Free Systems: Implementation and Debugging of a System of Biotransformations. Biotechnol. Bioeng. 2010, 106 (3), 376–389. 10.1002/bit.22666. [DOI] [PubMed] [Google Scholar]
- Garcia D. C.; Mohr B. P.; Dovgan J. T.; Hurst G. B.; Standaert R. F.; Doktycz M. J.. Elucidating the Potential of Crude Cell Extracts for Producing Pyruvate from Glucose. Synth. Biol. 2018, 3 ( (1), ). 10.1093/synbio/ysy006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dinglasan J. L. N.; Doktycz M. J.. Rewiring Cell-Free Metabolic Flux in E. Coli Lysates Using a Block—Push—Pull Approach. Synth. Biol. 2023, 8 ( (1), ). 10.1093/synbio/ysad007. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kay J. E.; Jewett M. C. Lysate of Engineered Escherichia Coli Supports High-Level Conversion of Glucose to 2,3-Butanediol. Metab. Eng. 2015, 32, 133–142. 10.1016/j.ymben.2015.09.015. [DOI] [PubMed] [Google Scholar]
- Rasor B. J.; Yi X.; Brown H.; Alper H. S.; Jewett M. C. An Integrated in Vivo/in Vitro Framework to Enhance Cell-Free Biosynthesis with Metabolically Rewired Yeast Extracts. Nat. Commun. 2021, 12 (1), 5139. 10.1038/s41467-021-25233-y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kay J. E.; Jewett M. C. A Cell-Free System for Production of 2,3-Butanediol Is Robust to Growth-Toxic Compounds. Metab. Eng. Commun. 2020, 10, e00114 10.1016/j.mec.2019.e00114. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Grubbe W. S.; Rasor B. J.; Krüger A.; Jewett M. C.; Karim A. S. Cell-Free Styrene Biosynthesis at High Titers. Metab. Eng. 2020, 61, 89–95. 10.1016/j.ymben.2020.05.009. [DOI] [PubMed] [Google Scholar]
- Karim A. S.; Rasor B. J.; Jewett M. C.. Enhancing Control of Cell-Free Metabolism through PH Modulation. Synth. Biol. 2020, 5 ( (1), ). 10.1093/synbio/ysz027. [DOI] [Google Scholar]
- Zhang P.; Wang J.; Ding X.; Lin J.; Jiang H.; Zhou H.; Lu Y. Exploration of the Tolerance Ability of a Cell-Free Biosynthesis System to Toxic Substances. Appl. Biochem. Biotechnol. 2019, 189 (4), 1096–1107. 10.1007/s12010-019-03039-5. [DOI] [PubMed] [Google Scholar]
- Keller M. W.; Schut G. J.; Lipscomb G. L.; Menon A. L.; Iwuchukwu I. J.; Leuko T. T.; Thorgersen M. P.; Nixon W. J.; Hawkins A. S.; Kelly R. M.; Adams M. W. W. Exploiting Microbial Hyperthermophilicity to Produce an Industrial Chemical, Using Hydrogen and Carbon Dioxide. Proc. Natl. Acad. Sci. U. S. A. 2013, 110 (15), 5840–5845. 10.1073/pnas.1222607110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hays S. G.; Yan L. L. W.; Silver P. A.; Ducat D. C. Synthetic Photosynthetic Consortia Define Interactions Leading to Robustness and Photoproduction. J. Biol. Eng. 2017, 11 (1), 4. 10.1186/s13036-017-0048-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jia X.; Liu C.; Song H.; Ding M.; Du J.; Ma Q.; Yuan Y. Design, Analysis and Application of Synthetic Microbial Consortia. Synth. Syst. Biotechnol. 2016, 1 (2), 109–117. 10.1016/j.synbio.2016.02.001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Panthu B.; Ohlmann T.; Perrier J.; Schlattner U.; Jalinot P.; Elena-Herrmann B.; Rautureau G. J. P. Cell-Free Protein Synthesis Enhancement from Real-Time NMR Metabolite Kinetics: Redirecting Energy Fluxes in Hybrid RRL Systems. ACS Synth. Biol. 2018, 7 (1), 218–226. 10.1021/acssynbio.7b00280. [DOI] [PubMed] [Google Scholar]
- Panthu B.; Décimo D.; Balvay L.; Ohlmann T. In Vitro Translation in a Hybrid Cell Free Lysate with Exogenous Cellular Ribosomes. Biochem. J. 2015, 467 (3), 387–398. 10.1042/BJ20141498. [DOI] [PubMed] [Google Scholar]
- Yi T.; Lim H. J.; Lee S. J.; Lee K.-H.; Kim D.-M. Synthesis of (R,R)-2,3-Butanediol from Starch in a Hybrid Cell-Free Reaction System. J. Ind. Eng. Chem. 2018, 67, 231. 10.1016/j.jiec.2018.06.033. [DOI] [Google Scholar]
- Bundy B. C.; Hunt J. P.; Jewett M. C.; Swartz J. R.; Wood D. W.; Frey D. D.; Rao G. Cell-Free Biomanufacturing. Curr. Opin. Chem. Eng. 2018, 22, 177–183. 10.1016/j.coche.2018.10.003. [DOI] [Google Scholar]
- Dudley Q. M.; Anderson K. C.; Jewett M. C. Cell-Free Mixing of Escherichia Coli Crude Extracts to Prototype and Rationally Engineer High-Titer Mevalonate Synthesis. ACS Synth. Biol. 2016, 5 (12), 1578–1588. 10.1021/acssynbio.6b00154. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Garcia D. C.; Dinglasan J. L. N.; Shrestha H.; Abraham P. E.; Hettich R. L.; Doktycz M. J. A Lysate Proteome Engineering Strategy for Enhancing Cell-Free Metabolite Production. Metab. Eng. Commun. 2021, 12, e00162 10.1016/j.mec.2021.e00162. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li X.; Guo D.; Cheng Y.; Zhu F.; Deng Z.; Liu T. Overproduction of Fatty Acids in Engineered Saccharomyces Cerevisiae. Biotechnol. Bioeng. 2014, 111 (9), 1841–1852. 10.1002/bit.25239. [DOI] [PubMed] [Google Scholar]
- Park C.-G.; Kwon M.-A.; Song J.-K.; Kim D.-M. Cell-Free Synthesis and Multifold Screening of Candida Antarctica Lipase B (CalB) Variants after Combinatorial Mutagenesis of Hot Spots. Biotechnol. Prog. 2011, 27 (1), 47–53. 10.1002/btpr.532. [DOI] [PubMed] [Google Scholar]
- Bhadra S.; Pothukuchy A.; Shroff R.; Cole A. W.; Byrom M.; Ellefson J. W.; Gollihar J. D.; Ellington A. D. Cellular Reagents for Diagnostics and Synthetic Biology. PLoS One 2018, 13 (8), e0201681 10.1371/journal.pone.0201681. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Bhadra S.; Nguyen V.; Torres J.-A.; Kar S.; Fadanka S.; Gandini C.; Akligoh H.; Paik I.; Maranhao A. C.; Molloy J.; Ellington A. D. Producing Molecular Biology Reagents without Purification. PLoS One 2021, 16 (6), e0252507 10.1371/journal.pone.0252507. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Garcia D. C.; Cheng X.; Land M. L.; Standaert R. F.; Morrell-Falvey J. L.; Doktycz M. J. Computationally Guided Discovery and Experimental Validation of Indole-3-Acetic Acid Synthesis Pathways. ACS Chem. Biol. 2019, 14 (12), 2867–2875. 10.1021/acschembio.9b00725. [DOI] [PubMed] [Google Scholar]
- Dudley Q. M.; Nash C. J.; Jewett M. C. Cell-Free Biosynthesis of Limonene Using Enzyme-Enriched Escherichia Coli Lysates. Synth. Biol. 2019, 4 (1), ysz003. 10.1093/synbio/ysz003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Niu F.-X.; Huang Y.-B.; Shen Y.-P.; Ji L.-N.; Liu J.-Z. Enhanced Production of Pinene by Using a Cell-Free System with Modular Cocatalysis. J. Agric. Food Chem. 2020, 68 (7), 2139–2145. 10.1021/acs.jafc.9b07830. [DOI] [PubMed] [Google Scholar]
- Niu F.-X.; Yan Z.-B.; Huang Y.-B.; Liu J.-Z. Cell-Free Biosynthesis of Chlorogenic Acid Using a Mixture of Chassis Cell Extracts and Purified Spy-Cyclized Enzymes. J. Agric. Food Chem. 2021, 69 (28), 7938–7947. 10.1021/acs.jafc.1c03309. [DOI] [PubMed] [Google Scholar]
- Karim A. S.; Jewett M. C. Cell-Free Synthetic Biology for Pathway Prototyping. Methods Enzymol. 2018, 608, 31–57. 10.1016/bs.mie.2018.04.029. [DOI] [PubMed] [Google Scholar]
- Rasor B. J.; Vögeli B.; Jewett M. C.; Karim A. S. Cell-Free Protein Synthesis for High-Throughput Biosynthetic Pathway Prototyping. Methods Mol. Biol. 2022, 2433, 199–215. 10.1007/978-1-0716-1998-8_12. [DOI] [PubMed] [Google Scholar]
- Karim A. S.; Dudley Q. M.; Juminaga A.; Yuan Y.; Crowe S. A.; Heggestad J. T.; Garg S.; Abdalla T.; Grubbe W. S.; Rasor B. J.; Coar D. N.; Torculas M.; Krein M.; Liew F. E.; Quattlebaum A.; Jensen R. O.; Stuart J. A.; Simpson S. D.; Köpke M.; Jewett M. C. In Vitro Prototyping and Rapid Optimization of Biosynthetic Enzymes for Cell Design. Nat. Chem. Biol. 2020, 16 (8), 912–919. 10.1038/s41589-020-0559-0. [DOI] [PubMed] [Google Scholar]
- Dudley Q. M.; Karim A. S.; Nash C. J.; Jewett M. C. In Vitro Prototyping of Limonene Biosynthesis Using Cell-Free Protein Synthesis. Metab. Eng. 2020, 61, 251–260. 10.1016/j.ymben.2020.05.006. [DOI] [PubMed] [Google Scholar]
- Liew F. E.; Nogle R.; Abdalla T.; Rasor B. J.; Canter C.; Jensen R. O.; Wang L.; Strutz J.; Chirania P.; De Tissera S.; Mueller A. P.; Ruan Z.; Gao A.; Tran L.; Engle N. L.; Bromley J. C.; Daniell J.; Conrado R.; Tschaplinski T. J.; Giannone R. J.; Hettich R. L.; Karim A. S.; Simpson S. D.; Brown S. D.; Leang C.; Jewett M. C.; Köpke M. Carbon-Negative Production of Acetone and Isopropanol by Gas Fermentation at Industrial Pilot Scale. Nat. Biotechnol. 2022, 40 (3), 335–344. 10.1038/s41587-021-01195-w. [DOI] [PubMed] [Google Scholar]
- Kelwick R.; Ricci L.; Chee S. M.; Bell D.; Webb A. J.; Freemont P. S.. Cell-Free Prototyping Strategies for Enhancing the Sustainable Production of Polyhydroxyalkanoates Bioplastics. Synth. Biol. 2018, 3 ( (1), ). 10.1093/synbio/ysy016. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Zhuang L.; Huang S.; Liu W.-Q.; Karim A. S.; Jewett M. C.; Li J. Total in Vitro Biosynthesis of the Nonribosomal Macrolactone Peptide Valinomycin. Metab. Eng. 2020, 60, 37–44. 10.1016/j.ymben.2020.03.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Khatri Y.; Hohlman R. M.; Mendoza J.; Li S.; Lowell A. N.; Asahara H.; Sherman D. H. Multicomponent Microscale Biosynthesis of Unnatural Cyanobacterial Indole Alkaloids. ACS Synth. Biol. 2020, 9 (6), 1349–1360. 10.1021/acssynbio.0c00038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Claassens N. J.; Burgener S.; Vögeli B.; Erb T. J.; Bar-Even A. A Critical Comparison of Cellular and Cell-Free Bioproduction Systems. Curr. Opin. Biotechnol. 2019, 60, 221–229. 10.1016/j.copbio.2019.05.003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Swartz J. R. Expanding Biological Applications Using Cell-Free Metabolic Engineering: An Overview. Metab. Eng. 2018, 50, 156–172. 10.1016/j.ymben.2018.09.011. [DOI] [PubMed] [Google Scholar]
- Xu C.; Hu S.; Chen X. Artificial Cells: From Basic Science to Applications. Mater. Today (Kidlington) 2016, 19 (9), 516–532. 10.1016/j.mattod.2016.02.020. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kai L.; Schwille P. Cell-Free Protein Synthesis and Its Perspectives for Assembling Cells from the Bottom-Up. Adv. Biosyst. 2019, 3 (6), e1800322 10.1002/adbi.201800322. [DOI] [PubMed] [Google Scholar]
- Rideau E.; Dimova R.; Schwille P.; Wurm F. R.; Landfester K. Liposomes and Polymersomes: A Comparative Review towards Cell Mimicking. Chem. Soc. Rev. 2018, 47 (23), 8572–8610. 10.1039/C8CS00162F. [DOI] [PubMed] [Google Scholar]
- Hindley J. W.; Law R. V.; Ces O.. Membrane Functionalization in Artificial Cell Engineering. SN Appl. Sci. 2020, 2 ( (4), ). 10.1007/s42452-020-2357-4. [DOI] [Google Scholar]
- Stano P.; Carrara P.; Kuruma Y.; Pereira de Souza T.; Luisi P. L. Compartmentalized Reactions as a Case of Soft-Matter Biotechnology: Synthesis of Proteins and Nucleic Acids inside Lipid Vesicles. J. Mater. Chem. 2011, 21 (47), 18887. 10.1039/c1jm12298c. [DOI] [Google Scholar]
- Noireaux V.; Maeda Y. T.; Libchaber A. Development of an Artificial Cell, from Self-Organization to Computation and Self-Reproduction. Proc. Natl. Acad. Sci. U. S. A. 2011, 108 (9), 3473–3480. 10.1073/pnas.1017075108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kaczanowska M.; Rydén-Aulin M. Ribosome Biogenesis and the Translation Process in Escherichia Coli. Microbiol. Mol. Biol. Rev. 2007, 71 (3), 477–494. 10.1128/MMBR.00013-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Traub P.; Nomura M. Structure and Function of E. Coli Ribosomes. V. Reconstitution of Functionally Active 30S Ribosomal Particles from RNA and Proteins. Proc. Natl. Acad. Sci. U. S. A. 1968, 59 (3), 777–784. 10.1073/pnas.59.3.777. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Culver G. M.; Noller H. F. Efficient Reconstitution of Functional Escherichia Coli 30S Ribosomal Subunits from a Complete Set of Recombinant Small Subunit Ribosomal Proteins. RNA 1999, 5 (6), 832–843. 10.1017/S1355838299990714. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Li J.; Wassie B.; Church G. M. Physiological Assembly of Functionally Active 30S Ribosomal Subunits Fromin Vitrosynthesized Parts. bioRxiv 2017, 10.1101/137745. [DOI] [Google Scholar]
- Murase Y.; Nakanishi H.; Tsuji G.; Sunami T.; Ichihashi N. In Vitro Evolution of Unmodified 16S RRNA for Simple Ribosome Reconstitution. ACS Synth. Biol. 2018, 7 (2), 576–583. 10.1021/acssynbio.7b00333. [DOI] [PubMed] [Google Scholar]
- Shimojo M.; Amikura K.; Masuda K.; Kanamori T.; Ueda T.; Shimizu Y. In Vitro Reconstitution of Functional Small Ribosomal Subunit Assembly for Comprehensive Analysis of Ribosomal Elements in E. Coli. Communications Biology 2020, 3 (1), 142–142. 10.1038/s42003-020-0874-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Tamaru D.; Amikura K.; Shimizu Y.; Nierhaus K. H.; Ueda T. Reconstitution of 30S Ribosomal Subunits in Vitro Using Ribosome Biogenesis Factors. RNA 2018, 24 (11), 1512–1519. 10.1261/rna.065615.118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Maki J. A.; Culver G. M. Recent Developments in Factor-Facilitated Ribosome Assembly. Methods 2005, 36 (3), 313–320. 10.1016/j.ymeth.2005.04.008. [DOI] [PubMed] [Google Scholar]
- Nierhaus K. H.; Dohme F. Total Reconstitution of Functionally Active 50S Ribosomal Subunits from Escherichia Coli. Proc. Natl. Acad. Sci. U. S. A. 1974, 71 (12), 4713–4717. 10.1073/pnas.71.12.4713. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Röhl R.; Nierhaus K. H. Assembly Map of the Large Subunit (50S) of Escherichia Coli Ribosomes. Proc. Natl. Acad. Sci. U. S. A. 1982, 79 (3), 729–733. 10.1073/pnas.79.3.729. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Aoyama R.; Masuda K.; Shimojo M.; Kanamori T.; Ueda T.; Shimizu Y. In Vitro Reconstitution of the Escherichia Coli 70S Ribosome with a Full Set of Recombinant Ribosomal Proteins. J. Biochem. 2022, 171 (2), 227–237. 10.1093/jb/mvab121. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Green R.; Noller H. F. In Vitro Complementation Analysis Localizes 23S RRNA Posttranscriptional Modifications That Are Required for Escherichia Coli 50S Ribosomal Subunit Assembly and Function. RNA 1996, 2 (10), 1011–1021. [PMC free article] [PubMed] [Google Scholar]
- Semrad K.; Green R. Osmolytes Stimulate the Reconstitution of Functional 50S Ribosomes from in Vitro Transcripts of Escherichia Coli 23S RRNA. RNA 2002, 8 (4), 401–411. 10.1017/S1355838202029722. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jewett M. C.; Fritz B. R.; Timmerman L. E.; Church G. M. In Vitro Integration of Ribosomal RNA Synthesis, Ribosome Assembly, and Translation. Mol. Syst. Biol. 2013, 9 (1), 1–8. 10.1038/msb.2013.31. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu Y.; Fritz B. R.; Anderson M. J.; Schoborg J. A.; Jewett M. C. Characterizing and Alleviating Substrate Limitations for Improved in Vitro Ribosome Construction. ACS Synth. Biol. 2015, 4 (4), 454–462. 10.1021/sb5002467. [DOI] [PubMed] [Google Scholar]
- Fritz B. R.; Jewett M. C. The Impact of Transcriptional Tuning on in Vitro Integrated RRNA Transcription and Ribosome Construction. Nucleic Acids Res. 2014, 42 (10), 6774–6785. 10.1093/nar/gku307. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Fritz B. R.; Jamil O. K.; Jewett M. C. Implications of Macromolecular Crowding and Reducing Conditions for in Vitro Ribosome Construction. Nucleic Acids Res. 2015, 43 (9), 4774–4784. 10.1093/nar/gkv329. [DOI] [PMC free article] [PubMed] [Google Scholar]
- d’Aquino A. E.; Azim T.; Aleksashin N. A.; Hockenberry A. J.; Krüger A.; Jewett M. C. Mutational Characterization and Mapping of the 70S Ribosome Active Site. Nucleic Acids Res. 2020, 48 (5), 2777–2789. 10.1093/nar/gkaa001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kofman C.; Watkins A. M.; Kim D. S.; Willi J. A.; Wooldredge A. C.; Karim A. S.; Das R.; Jewett M. C. Computationally-Guided Design and Selection of High Performing Ribosomal Active Site Mutants. Nucleic Acids Res. 2022, 50 (22), 13143–13154. 10.1093/nar/gkac1036. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Caschera F.; Lee J. W.; Ho K. K. Y.; Liu A. P.; Jewett M. C. Cell-Free Compartmentalized Protein Synthesis inside Double Emulsion Templated Liposomes with in Vitro Synthesized and Assembled Ribosomes. Chem. Commun. (Camb.) 2016, 52 (31), 5467–5469. 10.1039/C6CC00223D. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kofman C.; Willi J. A.; Karim A. S.; Jewett M. C. Ribosome Pool Engineering Increases Protein Biosynthesis Yields. ACS Cent. Sci. 2024, 10 (4), 871–881. 10.1021/acscentsci.3c01413. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Ludgate L.; Liu K.; Luckenbaugh L.; Streck N.; Eng S.; Voitenleitner C.; Delaney W. E. 4th; Hu J. Cell-Free Hepatitis B Virus Capsid Assembly Dependent on the Core Protein C-Terminal Domain and Regulated by Phosphorylation. J. Virol. 2016, 90 (12), 5830–5844. 10.1128/JVI.00394-16. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Klein K. C.; Dellos S. R.; Lingappa J. R. Identification of Residues in the Hepatitis C Virus Core Protein That Are Critical for Capsid Assembly in a Cell-Free System. J. Virol. 2005, 79 (11), 6814–6826. 10.1128/JVI.79.11.6814-6826.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Spice A. J.; Aw R.; Bracewell D. G.; Polizzi K. M. Synthesis and Assembly of Hepatitis B Virus-like Particles in a Pichia Pastoris Cell-Free System. Front. Bioeng. Biotechnol. 2020, 8, 72. 10.3389/fbioe.2020.00072. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Newcomb W. W.; Homa F. L.; Thomsen D. R.; Ye Z.; Brown J. C. Cell-Free Assembly of the Herpes Simplex Virus Capsid. J. Virol. 1994, 68 (9), 6059–6063. 10.1128/jvi.68.9.6059-6063.1994. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Newcomb W. W.; Homa F. L.; Thomsen D. R.; Booy F. P.; Trus B. L.; Steven A. C.; Spencer J. V.; Brown J. C. Assembly of the Herpes Simplex Virus Capsid: Characterization of Intermediates Observed during Cell-Free Capsid Formation. J. Mol. Biol. 1996, 263 (3), 432–446. 10.1006/jmbi.1996.0587. [DOI] [PubMed] [Google Scholar]
- Patton J. T.; Davis N. L.; Wertz G. W. Cell-Free Synthesis and Assembly of Vesicular Stomatitis Virus Nucleocapsids. J. Virol. 1983, 45 (1), 155–164. 10.1128/jvi.45.1.155-164.1983. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Varasteh Moradi S.; Gagoski D.; Mureev S.; Walden P.; McMahon K.-A.; Parton R. G.; Johnston W. A.; Alexandrov K. Mapping Interactions among Cell-Free Expressed Zika Virus Proteins. J. Proteome Res. 2020, 19 (4), 1522–1532. 10.1021/acs.jproteome.9b00771. [DOI] [PubMed] [Google Scholar]
- Bundy B. C.; Franciszkowicz M. J.; Swartz J. R. Escherichia Coli-Based Cell-Free Synthesis of Virus-like Particles. Biotechnol. Bioeng. 2008, 100 (1), 28–37. 10.1002/bit.21716. [DOI] [PubMed] [Google Scholar]
- Pogany J.; Nagy P. D. Authentic Replication and Recombination of Tomato Bushy Stunt Virus RNA in a Cell-Free Extract from Yeast. J. Virol. 2008, 82 (12), 5967–5980. 10.1128/JVI.02737-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Molla A.; Paul A. V.; Wimmer E. Cell-Free, de Novo Synthesis of Poliovirus. Science 1991, 254 (5038), 1647–1651. 10.1126/science.1661029. [DOI] [PubMed] [Google Scholar]
- Barton D. J.; Flanegan J. B. Coupled Translation and Replication of Poliovirus RNA in Vitro: Synthesis of Functional 3D Polymerase and Infectious Virus. J. Virol. 1993, 67 (2), 822–831. 10.1128/jvi.67.2.822-831.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Cello J.; Paul A. V.; Wimmer E. Chemical Synthesis of Poliovirus CDNA: Generation of Infectious Virus in the Absence of Natural Template. Science 2002, 297 (5583), 1016–1018. 10.1126/science.1072266. [DOI] [PubMed] [Google Scholar]
- Cuconati A.; Molla A.; Wimmer E. Brefeldin A Inhibits Cell-Free, de Novo Synthesis of Poliovirus. J. Virol. 1998, 72 (8), 6456–6464. 10.1128/JVI.72.8.6456-6464.1998. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Svitkin Y. V.; Sonenberg N. Cell-Free Synthesis of Encephalomyocarditis Virus. J. Virol. 2003, 77 (11), 6551–6555. 10.1128/JVI.77.11.6551-6555.2003. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kobayashi T.; Mikami S.; Yokoyama S.; Imataka H. An Improved Cell-Free System for Picornavirus Synthesis. J. Virol. Methods 2007, 142 (1–2), 182–188. 10.1016/j.jviromet.2007.01.026. [DOI] [PubMed] [Google Scholar]
- Kobayashi T.; Nakamura Y.; Mikami S.; Masutani M.; Machida K.; Imataka H. Synthesis of Encephalomyocarditis Virus in a Cell-Free System: From DNA to RNA Virus in One Tube. Biotechnol. Lett. 2012, 34 (1), 67–73. 10.1007/s10529-011-0744-z. [DOI] [PubMed] [Google Scholar]
- Shin J.; Jardine P.; Noireaux V. Genome Replication, Synthesis, and Assembly of the Bacteriophage T7 in a Single Cell-Free Reaction. ACS Synth. Biol. 2012, 1 (9), 408–413. 10.1021/sb300049p. [DOI] [PubMed] [Google Scholar]
- Rustad M.; Eastlund A.; Jardine P.; Noireaux V.. Cell-Free TXTL Synthesis of Infectious Bacteriophage T4 in a Single Test Tube Reaction. Synth. Biol. 2018, 3 ( (1), ). 10.1093/synbio/ysy002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Emslander Q.; Vogele K.; Braun P.; Stender J.; Willy C.; Joppich M.; Hammerl J. A.; Abele M.; Meng C.; Pichlmair A.; Ludwig C.; Bugert J. J.; Simmel F. C.; Westmeyer G. G. Cell-Free Production of Personalized Therapeutic Phages Targeting Multidrug-Resistant Bacteria. Cell Chem. Biol. 2022, 29 (9), 1434–1445. 10.1016/j.chembiol.2022.06.003. [DOI] [PubMed] [Google Scholar]
- Levrier A.; Karpathakis I.; Nash B.; Bowden S. D.; Lindner A. B.; Noireaux V.. PHEIGES: All-Cell-Free Phage Synthesis and Selection from Engineered Genomes. Nat. Commun. 2024, 15 ( (1), ). 10.1038/s41467-024-46585-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liyanagedera S. B. W.; Williams J.; Wheatley J. P.; Biketova A. Y.; Hasan M.; Sagona A. P.; Purdy K. J.; Puxty R. J.; Feher T.; Kulkarni V. SpyPhage: A Cell-Free TXTL Platform for Rapid Engineering of Targeted Phage Therapies. ACS Synth. Biol. 2022, 11 (10), 3330–3342. 10.1021/acssynbio.2c00244. [DOI] [PubMed] [Google Scholar]
- Dion M. B.; Oechslin F.; Moineau S. Phage Diversity, Genomics and Phylogeny. Nat. Rev. Microbiol. 2020, 18 (3), 125–138. 10.1038/s41579-019-0311-5. [DOI] [PubMed] [Google Scholar]
- Koonin E. V.; Dolja V. V.; Krupovic M. Origins and Evolution of Viruses of Eukaryotes: The Ultimate Modularity. Virology 2015, 479–480, 2–25. 10.1016/j.virol.2015.02.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wimmer E. The Test-Tube Synthesis of a Chemical Called Poliovirus. The Simple Synthesis of a Virus Has Far-Reaching Societal Implications. EMBO Rep. 2006, 7, S3–S9. 10.1038/sj.embor.7400728. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Jiang W.; Wu Z.; Gao Z.; Wan M.; Zhou M.; Mao C.; Shen J. Artificial Cells: Past, Present and Future. ACS Nano 2022, 16 (10), 15705–15733. 10.1021/acsnano.2c06104. [DOI] [PubMed] [Google Scholar]
- Krinsky N.; Kaduri M.; Zinger A.; Shainsky-Roitman J.; Goldfeder M.; Benhar I.; Hershkovitz D.; Schroeder A. Synthetic Cells Synthesize Therapeutic Proteins inside Tumors. Adv. Healthc. Mater. 2018, 7 (9), 1701163. 10.1002/adhm.201701163. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Nuti N.; Rottmann P.; Stucki A.; Koch P.; Panke S.; Dittrich P. S. A Multiplexed Cell-Free Assay to Screen for Antimicrobial Peptides in Double Emulsion Droplets. Angew. Chem., Int. Ed. Engl. 2022, 61 (13), e202114632 10.1002/anie.202114632. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Van Raad D.; Huber T. In Vitro Protein Synthesis in Semipermeable Artificial Cells. ACS Synth. Biol. 2021, 10 (5), 1237–1244. 10.1021/acssynbio.1c00044. [DOI] [PubMed] [Google Scholar]
- Maeda Y. T.; Nakadai T.; Shin J.; Uryu K.; Noireaux V.; Libchaber A. Assembly of MreB Filaments on Liposome Membranes: A Synthetic Biology Approach. ACS Synth. Biol. 2012, 1 (2), 53–59. 10.1021/sb200003v. [DOI] [PubMed] [Google Scholar]
- Garenne D.; Libchaber A.; Noireaux V. Membrane Molecular Crowding Enhances MreB Polymerization to Shape Synthetic Cells from Spheres to Rods. Proc. Natl. Acad. Sci. U. S. A. 2020, 117 (4), 1902–1909. 10.1073/pnas.1914656117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Kattan J.; Doerr A.; Dogterom M.; Danelon C. Shaping Liposomes by Cell-Free Expressed Bacterial Microtubules. ACS Synth. Biol. 2021, 10 (10), 2447–2455. 10.1021/acssynbio.1c00278. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lee K. Y.; Park S. J.; Lee K. A.; Kim S. H.; Kim H.; Meroz Y.; Mahadevan L.; Jung K. H.; Ahn T. K.; Parker K. K.; Shin K. Photosynthetic Artificial Organelles Sustain and Control ATP-Dependent Reactions in a Protocellular System. Nat. Biotechnol. 2018, 36 (6), 530–535. 10.1038/nbt.4140. [DOI] [PubMed] [Google Scholar]
- Godino E.; López J. N.; Foschepoth D.; Cleij C.; Doerr A.; Castellà C. F.; Danelon C. De Novo Synthesized Min Proteins Drive Oscillatory Liposome Deformation and Regulate FtsA-FtsZ Cytoskeletal Patterns. Nat. Commun. 2019, 10 (1), 4969. 10.1038/s41467-019-12932-w. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Furusato T.; Horie F.; Matsubayashi H. T.; Amikura K.; Kuruma Y.; Ueda T. De Novo Synthesis of Basal Bacterial Cell Division Proteins FtsZ, FtsA, and ZipA inside Giant Vesicles. ACS Synth. Biol. 2018, 7 (4), 953–961. 10.1021/acssynbio.7b00350. [DOI] [PubMed] [Google Scholar]
- Godino E.; López J. N.; Zarguit I.; Doerr A.; Jimenez M.; Rivas G.; Danelon C. Cell-Free Biogenesis of Bacterial Division Proto-Rings That Can Constrict Liposomes. Commun. Biol. 2020, 3 (1), 539. 10.1038/s42003-020-01258-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Rivas G.; Vogel S. K.; Schwille P. Reconstitution of Cytoskeletal Protein Assemblies for Large-Scale Membrane Transformation. Curr. Opin. Chem. Biol. 2014, 22, 18–26. 10.1016/j.cbpa.2014.07.018. [DOI] [PubMed] [Google Scholar]
- Boyd M. A.; Kamat N. P. Designing Artificial Cells towards a New Generation of Biosensors. Trends Biotechnol. 2021, 39 (9), 927–939. 10.1016/j.tibtech.2020.12.002. [DOI] [PubMed] [Google Scholar]
- Dubuc E.; Pieters P. A.; van der Linden A. J.; van Hest J. C.; Huck W. T.; de Greef T. F. Cell-Free Microcompartmentalised Transcription-Translation for the Prototyping of Synthetic Communication Networks. Curr. Opin. Biotechnol. 2019, 58, 72–80. 10.1016/j.copbio.2018.10.006. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Smith J. M.; Chowdhry R.; Booth M. J. Controlling Synthetic Cell-Cell Communication. Front. Mol. Biosci. 2022, 8, 809945. 10.3389/fmolb.2021.809945. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Hilburger C. E.; Jacobs M. L.; Lewis K. R.; Peruzzi J. A.; Kamat N. P. Controlling Secretion in Artificial Cells with a Membrane AND Gate. ACS Synth. Biol. 2019, 8 (6), 1224–1230. 10.1021/acssynbio.8b00435. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Dupin A.; Simmel F. C. Signalling and Differentiation in Emulsion-Based Multi-Compartmentalized in Vitro Gene Circuits. Nat. Chem. 2019, 11 (1), 32–39. 10.1038/s41557-018-0174-9. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Gaut N. J.; Gomez-Garcia J.; Heili J. M.; Cash B.; Han Q.; Engelhart A. E.; Adamala K. P. Programmable Fusion and Differentiation of Synthetic Minimal Cells. ACS Synth. Biol. 2022, 11 (2), 855–866. 10.1021/acssynbio.1c00519. [DOI] [PubMed] [Google Scholar]
- Peruzzi J. A.; Jacobs M. L.; Vu T. Q.; Wang K. S.; Kamat N. P. Barcoding Biological Reactions with DNA-functionalized Vesicles. Angew. Chem. Weinheim Bergstr. Ger. 2019, 131 (51), 18856–18863. 10.1002/ange.201911544. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Elani Y. Interfacing Living and Synthetic Cells as an Emerging Frontier in Synthetic Biology. Angew. Chem. Weinheim Bergstr. Ger. 2021, 133 (11), 5662–5671. 10.1002/ange.202006941. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Schwarz-Schilling M.; Aufinger L.; Mückl A.; Simmel F. C. Chemical Communication between Bacteria and Cell-Free Gene Expression Systems within Linear Chains of Emulsion Droplets. Integr. Biol. (Camb.) 2016, 8 (4), 564–570. 10.1039/C5IB00301F. [DOI] [PubMed] [Google Scholar]
- Rampioni G.; D’Angelo F.; Messina M.; Zennaro A.; Kuruma Y.; Tofani D.; Leoni L.; Stano P. Synthetic Cells Produce a Quorum Sensing Chemical Signal Perceived ByPseudomonas Aeruginosa. Chem. Commun. (Camb.) 2018, 54 (17), 2090–2093. 10.1039/C7CC09678J. [DOI] [PubMed] [Google Scholar]
- Ding Y.; Contreras-Llano L. E.; Morris E.; Mao M.; Tan C. Minimizing Context Dependency of Gene Networks Using Artificial Cells. ACS Appl. Mater. Interfaces 2018, 10 (36), 30137–30146. 10.1021/acsami.8b10029. [DOI] [PubMed] [Google Scholar]
- Lentini R.; Martín N. Y.; Forlin M.; Belmonte L.; Fontana J.; Cornella M.; Martini L.; Tamburini S.; Bentley W. E.; Jousson O.; Mansy S. S. Two-Way Chemical Communication between Artificial and Natural Cells. ACS Cent. Sci. 2017, 3 (2), 117–123. 10.1021/acscentsci.6b00330. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Lentini R.; Santero S. P.; Chizzolini F.; Cecchi D.; Fontana J.; Marchioretto M.; Del Bianco C.; Terrell J. L.; Spencer A. C.; Martini L.; Forlin M.; Assfalg M.; Dalla Serra M.; Bentley W. E.; Mansy S. S. Integrating Artificial with Natural Cells to Translate Chemical Messages That Direct E. Coli Behaviour. Nat. Commun. 2014, 5 (1), 4012. 10.1038/ncomms5012. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Toparlak Ö. D.; Zasso J.; Bridi S.; Serra M. D.; Macchi P.; Conti L.; Baudet M.-L.; Mansy S. S.. Artificial Cells Drive Neural Differentiation. Sci. Adv. 2020, 6 ( (38), ). 10.1126/sciadv.abb4920. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Elani Y.; Trantidou T.; Wylie D.; Dekker L.; Polizzi K.; Law R. V.; Ces O.. Constructing Vesicle-Based Artificial Cells with Embedded Living Cells as Organelle-like Modules. Sci. Rep. 2018, 8 ( (1), ). 10.1038/s41598-018-22263-3. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Love C.; Steinkühler J.; Gonzales D. T.; Yandrapalli N.; Robinson T.; Dimova R.; Tang T.-Y. D. Reversible PH-Responsive Coacervate Formation in Lipid Vesicles Activates Dormant Enzymatic Reactions. Angew. Chem., Int. Ed. Engl. 2020, 59 (15), 5950–5957. 10.1002/anie.201914893. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Liu J.; Guo Z.; Li Y.; Liang J.; Xue J.; Xu J.; Whitelock J. M.; Xie L.; Kong B.; Liang K.. PH-gated Activation of Gene Transcription and Translation in Biocatalytic Metal-Organic Framework Artificial Cells. Adv. Nanobiomed Res. 2021, 1 ( (1), ). 10.1002/anbr.202000034. [DOI] [Google Scholar]
- Hindley J. W.; Zheleva D. G.; Elani Y.; Charalambous K.; Barter L. M. C.; Booth P. J.; Bevan C. L.; Law R. V.; Ces O. Building a Synthetic Mechanosensitive Signaling Pathway in Compartmentalized Artificial Cells. Proc. Natl. Acad. Sci. U. S. A. 2019, 116 (34), 16711–16716. 10.1073/pnas.1903500116. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Majumder S.; Garamella J.; Wang Y.-L.; DeNies M.; Noireaux V.; Liu A. P. Cell-Sized Mechanosensitive and Biosensing Compartment Programmed with DNA. Chem. Commun. (Camb.) 2017, 53 (53), 7349–7352. 10.1039/C7CC03455E. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Garamella J.; Majumder S.; Liu A. P.; Noireaux V. An Adaptive Synthetic Cell Based on Mechanosensing, Biosensing, and Inducible Gene Circuits. ACS Synth. Biol. 2019, 8 (8), 1913–1920. 10.1021/acssynbio.9b00204. [DOI] [PubMed] [Google Scholar]
- Strutt R.; Hindley J. W.; Gregg J.; Booth P. J.; Harling J. D.; Law R. V.; Friddin M. S.; Ces O. Activating Mechanosensitive Channels Embedded in Droplet Interface Bilayers Using Membrane Asymmetry. Chem. Sci. 2021, 12 (6), 2138–2145. 10.1039/D0SC03889J. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Manzer Z. A.; Ghosh S.; Jacobs M. L.; Krishnan S.; Zipfel W. R.; Piñeros M.; Kamat N. P.; Daniel S. Cell-Free Synthesis of a Transmembrane Mechanosensitive Channel Protein into a Hybrid-Supported Lipid Bilayer. ACS Appl. Bio Mater. 2021, 4 (4), 3101–3112. 10.1021/acsabm.0c01482. [DOI] [PubMed] [Google Scholar]
- Kleineberg C.; Wölfer C.; Abbasnia A.; Pischel D.; Bednarz C.; Ivanov I.; Heitkamp T.; Börsch M.; Sundmacher K.; Vidaković-Koch T. Light-Driven ATP Regeneration in Diblock/Grafted Hybrid Vesicles. Chembiochem 2020, 21 (15), 2149–2160. 10.1002/cbic.201900774. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Berhanu S.; Ueda T.; Kuruma Y. Artificial Photosynthetic Cell Producing Energy for Protein Synthesis. Nat. Commun. 2019, 10 (1), 1325. 10.1038/s41467-019-09147-4. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Altamura E.; Albanese P.; Marotta R.; Milano F.; Fiore M.; Trotta M.; Stano P.; Mavelli F. Chromatophores Efficiently Promote Light-Driven ATP Synthesis and DNA Transcription inside Hybrid Multicompartment Artificial Cells. Proc. Natl. Acad. Sci. U. S. A. 2021, 118 (7), e2012170118 10.1073/pnas.2012170118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Miller T. E.; Beneyton T.; Schwander T.; Diehl C.; Girault M.; McLean R.; Chotel T.; Claus P.; Cortina N. S.; Baret J.-C.; Erb T. J. Light-Powered CO2 Fixation in a Chloroplast Mimic with Natural and Synthetic Parts. Science 2020, 368 (6491), 649–654. 10.1126/science.aaz6802. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Libicher K.; Hornberger R.; Heymann M.; Mutschler H. In Vitro Self-Replication and Multicistronic Expression of Large Synthetic Genomes. Nat. Commun. 2020, 11 (1), 904. 10.1038/s41467-020-14694-2. [DOI] [PMC free article] [PubMed] [Google Scholar]
- van Nies P.; Westerlaken I.; Blanken D.; Salas M.; Mencía M.; Danelon C.. Self-Replication of DNA by Its Encoded Proteins in Liposome-Based Synthetic Cells. Nat. Commun. 2018, 9 ( (1), ). 10.1038/s41467-018-03926-1. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Doerr A.; Foschepoth D.; Forster A. C.; Danelon C. In Vitro Synthesis of 32 Translation-Factor Proteins from a Single Template Reveals Impaired Ribosomal Processivity. Sci. Rep. 2021, 11 (1), 1898. 10.1038/s41598-020-80827-8. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Libicher K.; Mutschler H. Probing Self-Regeneration of Essential Protein Factors Required for in Vitro Translation Activity by Serial Transfer. Chem. Commun. (Camb.) 2020, 56 (98), 15426–15429. 10.1039/D0CC06515C. [DOI] [PubMed] [Google Scholar]
- Lavickova B.; Laohakunakorn N.; Maerkl S. J. A Partially Self-Regenerating Synthetic Cell. Nat. Commun. 2020, 11 (1), 6340. 10.1038/s41467-020-20180-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Wei E.; Endy D. Experimental Tests of Functional Molecular Regeneration via a Standard Framework for Coordinating Synthetic Cell Building. bioRxiv 2021, 10.1101/2021.03.03.433818. [DOI] [Google Scholar]
- Li J.; Haas W.; Jackson K.; Kuru E.; Jewett M. C.; Fan Z. H.; Gygi S.; Church G. M. Cogenerating Synthetic Parts toward a Self-Replicating System. ACS Synth. Biol. 2017, 6 (7), 1327–1336. 10.1021/acssynbio.6b00342. [DOI] [PubMed] [Google Scholar]
- Bhattacharya A.; Brea R. J.; Niederholtmeyer H.; Devaraj N. K. A Minimal Biochemical Route towards de Novo Formation of Synthetic Phospholipid Membranes. Nat. Commun. 2019, 10 (1), 300. 10.1038/s41467-018-08174-x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Blanken D.; Foschepoth D.; Serrão A. C.; Danelon C. Genetically Controlled Membrane Synthesis in Liposomes. Nat. Commun. 2020, 11 (1), 4317. 10.1038/s41467-020-17863-5. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Garrett T. A.; Osmundson J.; Isaacson M.; Herrera J. Doing That Thing That Scientists Do: A Discovery-Driven Module on Protein Purification and Characterization for the Undergraduate Biochemistry Laboratory Classroom. Biochem. Mol. Biol. Educ. 2015, 43 (3), 145–153. 10.1002/bmb.20844. [DOI] [PubMed] [Google Scholar]
- Junker M. A Hands-on Classroom Simulation to Demonstrate Concepts in Enzyme Kinetics. J. Chem. Educ. 2010, 87 (3), 294–295. 10.1021/ed800082f. [DOI] [Google Scholar]
- Kelwick R.; Bowater L.; Yeoman K. H.; Bowater R. P. Promoting Microbiology Education through the IGEM Synthetic Biology Competition. FEMS Microbiol. Lett. 2015, 362 (16), fnv129. 10.1093/femsle/fnv129. [DOI] [PubMed] [Google Scholar]
- Kuldell N.; Hart K. M.; Bernstein R.; Ingram K.. BioBuilder; O’Reilly Media: Sebastopol, CA, 2015. [Google Scholar]
- 2023 CSHL Synthetic Biology Course. https://meetings.cshl.edu/courses.aspx?course=c-synbio&year=23 (accessed 2024-01-28).
- Zhang P.; Yang J.; Cho E.; Lu Y. Bringing Light into Cell-Free Expression. ACS Synth. Biol. 2020, 9 (8), 2144–2153. 10.1021/acssynbio.0c00211. [DOI] [PubMed] [Google Scholar]
- Collias D.; Marshall R.; Collins S. P.; Beisel C. L.; Noireaux V.. An Educational Module to Explore CRISPR Technologies with a Cell-Free Transcription-Translation System. Synth. Biol. 2019, 4 ( (1), ). 10.1093/synbio/ysz005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Stark J. C.; Huang A.; Hsu K. J.; Dubner R. S.; Forbrook J.; Marshalla S.; Rodriguez F.; Washington M.; Rybnicky G. A.; Nguyen P. Q.; Hasselbacher B.; Jabri R.; Kamran R.; Koralewski V.; Wightkin W.; Martinez T.; Jewett M. C. BioBits Health: Classroom Activities Exploring Engineering, Biology, and Human Health with Fluorescent Readouts. ACS Synth. Biol. 2019, 8 (5), 1001–1009. 10.1021/acssynbio.8b00381. [DOI] [PubMed] [Google Scholar]
- Rybnicky G. A.; Dixon R. A.; Kuhn R. M.; Karim A. S.; Jewett M. C. Development of a Freeze-Dried CRISPR-Cas12 Sensor for Detecting Wolbachia in the Secondary Science Classroom. ACS Synth. Biol. 2022, 11 (2), 835–842. 10.1021/acssynbio.1c00503. [DOI] [PubMed] [Google Scholar]
- Collins M. C.; Lau M. B.; Ma W.; Shen A.; Russa M. L.; Qi L. S. Biodesign of Frugal CRISPR Kits for Equitable and Accessible Education. bioRxiv 2023, 10.1101/2023.10.14.562372. [DOI] [Google Scholar]
- Hallinan J. S.; Wipat A.; Kitney R.; Woods S.; Taylor K.; Goñi-Moreno A. Future-proofing Synthetic Biology: Educating the next Generation. Eng. biol. 2019, 3 (2), 25–31. 10.1049/enb.2019.0001. [DOI] [Google Scholar]
- Williams L. C.; Gregorio N. E.; So B.; Kao W. Y.; Kiste A. L.; Patel P. A.; Watts K. R.; Oza J. P.. The Genetic Code Kit: An Open-Source Cell-Free Platform for Biochemical and Biotechnology Education. Front. Bioeng. Biotechnol. 2020, 8. 10.3389/fbioe.2020.00941. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Huang A.; Bryan B.; Kraves S.; Alvarez-Saavedra E.; Stark J. C. Implementing Hands-on Molecular and Synthetic Biology Education Using Cell-Free Technology. Methods Mol. Biol. 2022, 2433, 413–432. 10.1007/978-1-0716-1998-8_25. [DOI] [PubMed] [Google Scholar]
- Luo S.; Adam D.; Giaveri S.; Barthel S.; Cestellos-Blanco S.; Hege D.; Paczia N.; Castañeda-Losada L.; Klose M.; Arndt F.; Heider J.; Erb T. J. ATP Production from Electricity with a New-to-Nature Electrobiological Module. Joule 2023, 7 (8), 1745–1758. 10.1016/j.joule.2023.07.012. [DOI] [Google Scholar]
- Smith L. M.; Kelleher N. L. Proteoforms as the next Proteomics Currency. Science 2018, 359 (6380), 1106–1107. 10.1126/science.aat1884. [DOI] [PMC free article] [PubMed] [Google Scholar]
- Levine M. Z.; Gregorio N. E.; Jewett M. C.; Watts K. R.; Oza J. P. Escherichia Coli-Based Cell-Free Protein Synthesis: Protocols for a Robust, Flexible, and Accessible Platform Technology. J. Vis. Exp. 2019, 144, e58882. 10.3791/58882. [DOI] [PubMed] [Google Scholar]







