Targetron-Assisted Delivery of Exogenous DNA Sequences into Pseudomonas putida through CRISPR-Aided Counterselection
- Elena VelázquezElena VelázquezSystems and Synthetic Biology Department, Centro Nacional de Biotecnología (CNB-CSIC), Campus de Cantoblanco, Madrid 28049, SpainMore by Elena Velázquez
- ,
- Yamal Al-RamahiYamal Al-RamahiSystems and Synthetic Biology Department, Centro Nacional de Biotecnología (CNB-CSIC), Campus de Cantoblanco, Madrid 28049, SpainMore by Yamal Al-Ramahi
- ,
- Jonathan Tellechea-LuzardoJonathan Tellechea-LuzardoInterdisciplinary Computing and Complex Biosystems (ICOS) Research Group, Newcastle University, Newcastle Upon Tyne NE4 5TG, U.K.More by Jonathan Tellechea-Luzardo
- ,
- Natalio KrasnogorNatalio KrasnogorInterdisciplinary Computing and Complex Biosystems (ICOS) Research Group, Newcastle University, Newcastle Upon Tyne NE4 5TG, U.K.More by Natalio Krasnogor
- , and
- Víctor de Lorenzo*Víctor de Lorenzo*Email: [email protected]. Tel: 34-91 585 45 36. Fax: 34-91 585 45 06.Systems and Synthetic Biology Department, Centro Nacional de Biotecnología (CNB-CSIC), Campus de Cantoblanco, Madrid 28049, SpainMore by Víctor de Lorenzo
Abstract

Genome editing methods based on group II introns (known as targetron technology) have long been used as a gene knockout strategy in a wide range of organisms, in a fashion independent of homologous recombination. Yet, their utility as delivery systems has typically been suboptimal due to the reduced efficiency of insertion when carrying exogenous sequences. We show that this limitation can be tackled and targetrons can be adapted as a general tool in Gram-negative bacteria. To this end, a set of broad-host-range standardized vectors were designed for the conditional expression of the Ll.LtrB intron. After establishing the correct functionality of these plasmids in Escherichia coli and Pseudomonas putida, we created a library of Ll.LtrB variants carrying cargo DNA sequences of different lengths, to benchmark the capacity of intron-mediated delivery in these bacteria. Next, we combined CRISPR/Cas9-facilitated counterselection to increase the chances of finding genomic sites inserted with the thereby engineered introns. With these novel tools, we were able to insert exogenous sequences of up to 600 bp at specific genomic locations in wild-type P. putida KT2440 and its ΔrecA derivative. Finally, we applied this technology to successfully tag P. putida with an orthogonal short sequence barcode that acts as a unique identifier for tracking this microorganism in biotechnological settings. These results show the value of the targetron approach for the unrestricted delivery of small DNA fragments to precise locations in the genomes of Gram-negative bacteria, which will be useful for a suite of genome editing endeavors.
This publication is licensed under
License Summary*
You are free to share (copy and redistribute) this article in any medium or format and to adapt (remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share (copy and redistribute) this article in any medium or format and to adapt (remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share (copy and redistribute) this article in any medium or format and to adapt (remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
Results and Discussion
Engineering Broad-Host Expression of Intron Ll.LtrB
Figure 1

Figure 1. Diagram of targetrons and CRISPR/Cas9-mediated counterselection of insertions. The figure shows the technology developed in this article. Plasmid pSEVA-GIIi expresses the Ll.LtrB group II intron (empty or with cargo sequences cloned in the MluI site) and its IEP (LtrA) in the same transcriptional unit from the upstream promoter. The transcriptional unit is led by a retargeting region at 5′ (including exon 1 and the 5′ sequence of the Ll.LtrB intron), where three short sequences retrieved from the target gene (IBS, EBS1d, and EBS2) were engineered at given sites of the predicted transcript to secure its proper folding and retargeting (location of diagnostic primers is indicated). After transcription, the intronic RNA folds into a very conserved secondary structure and associates with LtrA to perform the splicing process from the exons. A lariat RNA is generated that remains attached to LtrA, forming a ribonucleoprotein (RNP) complex. This RNP scans DNA molecules until it finds the target site for retrohoming. Reverse splicing links covalently the intronic RNA to the sense strand of the DNA molecule, and the endonuclease (En) domain present in LtrA cleaves the antisense strand. Afterward, the retrotranscriptase (RT) domain of LtrA reverse transcribes the intronic RNA into DNA. The complete integration and synthesis of the cDNA is driven by repair mechanisms without the involvement of recombination. The incorporation of Cas9 complexes with gRNAs that recognize the insertion locus of Ll.LtrB causes the elimination of nonedited cells, as only those which incorporated the group II intron at the correct locus can survive (counterselection). IBS1: intron-binding site 1; EBS1d: exon-binding site 1-δ; EBS2: exon-binding site 2.
Figure 2

Figure 2. SEVA plasmids encoding the Ll.LtrB group II intron and T7 RNAP work in E. coli BL21DE3 and P. putida KT2440. (A) Delivery of the Ll.LtrB intron from plasmid pSEVA421-GIIi (Km) in E. coli BL21DE3. The Ll.LtrB intron was retargeted to insert in the antisense orientation into the locus 1063 of the lacZ gene so that insertions would disrupt this gene, giving rise to white colonies in the presence of X-gal. Since a RAM is placed inside Ll.LtrB, kanamycin resistance was used as a way to select for intron insertion mutants (plates to the right). (B) Graph shows the number of KmR CFU normalized to 109 viable cells and classified according to the displayed phenotype in the presence of X-gal (blue colonies: lacZ+, blue bars; white colonies: lacZ– (disrupted), white hatched bars) and, also, according to the presence or absence of IPTG induction. (C) Representative colony PCR reactions to determine the correct insertion of Ll.LtrB inside the lacZ gene. Only if Ll.LtrB retrohomes, a PCR amplicon of 720 bp is generated. Blue numbers correspond to blue colonies, and black numbers correspond to white colonies used as the template material for each reaction. (D–F) Delivery of Ll.LtrB from plasmid pSEVA421-GIIi-pyrF and with the help of pSEVA131-T7RNAP in P. putida KT2440. (D) Bar plot showing the frequency of 5FOAR CFU normalized to 109 viable cells after the insertion assay. CFU are classified according to the addition or absence of IPTG during the incubation period. (E) Genuine efficiency of the insertion of Ll.LtrB. The proportion of uracil auxotrophs detected from the 5FOAR population was used as a ratio to determine the abundance of Ll.LtrB insertions in the population. (F) 5FOA counterselection was used to isolate insertion mutants that were not able to grow without uracil supplemented to plates. Colonies resistant to 5FOA but that were able to grow without uracil were used as negative controls of insertion. Two different PCR reactions are shown: (top gel) one primer annealed inside Ll.LtrB and the second annealed in the pyrF gene so that an amplicon could be only generated after intron insertion. (Bottom gel) two primers flanking the insertion locus were used so that two amplicons could be generated. The smallest fragment (380 bp) corresponds to the WT sequence, and the biggest fragment (∼1500 bp) corresponds to the insertion. The same colonies were tested in both PCR reactions. All bar graphs show the mean values (bars), standard deviation (lines), and single values (dots) of two biological replicates. WT: wild-type; Mut: insertion mutant; + control: reaction with a colony with successful insertion from a previous experiment used as a template; H2O: control PCR with no template material; UraR: colonies growing in media without uracil (no auxotrophs); UraS: colonies not growing in media without uracil (auxotrophs).
Ll.LtrB Intron Retrohomes in P. putida KT2440
Streamlining Ll.LtrB Expression and Activity in a ΔrecA Derivative Strain of P. putida KT2440
Figure 3

Figure 3. Performance of pSEVA2311-GIIi-pyrF in P. putida KT2440 and its ΔrecA derivative strain. (A) pSEVA2311-GIIi-pyrF works in P. putida KT2440 to deliver the Ll.LtrB intron into the pyrF gene through 5FOA counterselection. Left panel: Bar plot showing the frequency of 5FOAR CFUs normalized to 109 viable cells of wild-type P. putida after the insertion assay resulting from induction with various cyclohexanone concentrations (0, 0.5, 1, and 5 mM). Right panel: Genuine efficiency of insertion of Ll.LtrB. The proportion of uracil auxotrophs detected from the 5FOAR population was used as a reference to determine the abundance of Ll.LtrB insertions in each population. (B) Colony PCR reaction that used primers flanking the insertion locus was used to determine Ll.LtrB retrohoming in each concentration of cyclohexanone tested (from 0 mM in the left part of gel to 5 mM in the right part of gel). The smallest fragment (380 bp) corresponds to the wild-type sequence, while the biggest one (1500 bp) corresponds to the intron insertion in the correct location. (C, D) Functioning of pSEVA2311-GIIi-pyrF in P. putida KT2440 ΔrecA. The same experiments and analyses were done to determine the frequencies and correctness of intron insertion in the recombination-deficient strain. The gel at the bottom of (D) shows a control PCR reaction to verify recA minus genotype of the tested cells. The frequency of 5FOAR CFUs and the efficiency of insertion of Ll.LtrB in P. putida ΔrecA were determined as before. All bar graphs show the mean values (bars), standard deviation (lines), and single values (dots) of at least three biological replicates. WT, wild-type; Mut, insertion mutant; KT2440, parental strain; + control: PCR of DNA from an intron-inserted colony (from a previous experiment) used as the template; H2O: control PCR with no template material.
Merging Ll.LtrB Action with a CRISPR/Cas9-Mediated Counterselection System
Figure 4

Figure 4. Assessing the size-restriction of intron-mediated delivery with CRISPR/Cas9 counterselection using luxC fragments as a cargo. (A) Schematic of the intron library generated with increasing fragment length as a cargo (from 150 up to 1050 bp) using as template the first gene of the luxCDABEG operon, luxC. The Ll.LtrB intron in pSEVA6511-GIIi (LuxN) is retargeted to insert between the nucleotides 165 and 166 of the pyrF ORF in the antisense orientation. Spacer pyrF1 recognizes the region after the insertion site (part of the recognition site is shown inside a red box). The complementary nucleotides to the PAM (5′GGG-3′) are highlighted in red and are disrupted upon intron insertion. (B) Ll.LtrB-mediated delivery of luxC fragments in P. putida KT2440 WT with CRISPR/Cas9 counterselection. Colony PCR reactions showing amplifications from colonies with Ll.LtrB::LuxØ and Ll.LtrB::Lux1 (top gel) and corresponding PCR reactions verifying the recA-plus phenotype (bottom gel). WT amplification for the recA gene is 2 kb long. (C) Ll.LtrB-mediated delivery of luxC fragments in P. putida KT2440 ΔrecA. Colony PCR reactions showing amplifications from colonies with Ll.LtrB::LuxØ to Ll.LtrB::Lux4 (top gel) and corresponding PCR reaction verifying the recA—genotype (bottom gel). Deletion of the recA gene gives an amplification of 1 kb. WT: wild-type, Mut: insertion mutant, LuxN: cargos including from LuxØ to Lux7, Ø: Ll.LtrB with no cargo, 1: Ll.LtrB with Lux1 as the cargo, 2: Ll.LtrB with Lux2 as the cargo; 3: Ll.LtrB with Lux3 as the cargo, and 4: Ll.LtrB with Lux4 as the cargo. P. putida KT2440 ΔrecA colonies with no inserted Ll.LtrB used as the negative control.
Figure 5

Figure 5. Intron insertion frequencies with CRISPR-Cas9 counterselection in wild-type and ΔrecAP. putida KT2440 confirmed through PCR. (A) Intron insertion frequency of each cargo inserted in the genome of WT P. putida KT2440 by Ll.LtrB using 5FOA; no counterselection (control plasmid pSEVA231-CRISPR) or CRISPR/Cas9-mediated counterselection (pSEVA231-C-pyrF1). Insertion of a cargo larger than Lux1 was not detected. (B) Similarly, for P. putida KT2440 ΔrecA, insertion of a cargo larger than Lux4 was not detected. (C) Combined efficiency of targetrons and CRISPR/Cas9 counterselection. The numbers of SmR KmR CFU obtained after transforming either pSEVA231-CRISPR (control) or pSEVA231-C-pyrF1 were normalized to 109 cells. The pSEVA231-CRISPR condition was set to 100%, and the percentage of cells with pSEVA231-C-pyrF1 was calculated accordingly. Finally, the ratio of positive Ll.LtrB insertions detected in the pSEVA231-C-pyrF1 condition was multiplied individually in each replicate. The mean (bars), single values (dots), and standard deviation (lines) of two or three replicates are shown. Ø: Ll.LtrB with no cargo; 1: Ll.LtrB with Lux1 as the cargo; 2: Ll.LtrB with Lux2 as the cargo; 3: Ll.LtrB with Lux3 as the cargo; and 4: Ll.LtrB with Lux4 as the cargo.
Application of Ll.LtrB for Barcoding Cells with Unique DNA Identifiers
Figure 6

Figure 6. Application of the Ll.LtrB group II intron for delivery of specific genetic barcodes to the genome of P. putida KT2440. Organization of pSEVA6511-GIIi (B3) variants is shown with the intron retargeted toward Locus 1 (x = 37s) or Locus 2 (x = 95a). Selection of the insertion loci for Ll.LtrB::B3 is in the vicinity of the Tn7-insertion site (black triangle). Two different insertion points (gray triangles) were chosen for the insertion list generated on the Clostron website, (24) and Ll.LtrB::B3 was retargeted to both sites accordingly. The recognition site in Locus 1 (green) is located in the sense strand, while Locus 2 (orange) is present in the antisense strand of the P. putida genome. Ll.LtrB::B3 insertion would generate two different genotypes depending on the locus being targeted in each case.
Figure 7

Figure 7. Delivery of Ll.LtrB::B3 containing a barcode in the P. putida KT2440 genome. A first pool PCR was set to detect successful Ll.LtrB::B3 insertions in either Locus 1 (green) or Locus 2 (orange). The top gel shows the amplification found with a pool PCR using primers flanking the insertion at Locus 2. The bottom gel shows the second PCR of individual colonies from the corresponding pool to find the barcoded clone. In this case, a primer annealing inside the barcode (pbarcode universal) and another annealing inside the PP5408 gene were used. WT: wild-type, Locus 1: 36,37s insertion site, Locus 2: 94,95a insertion site. PP5408 (green gene), glmS (orange gene).
Conclusions
Methods
Bacterial Strains and Media
Plasmid Construction
Retargeting of the Ll.LtrB Intron
Insertion of Exogenous Sequences Inside Ll.LtrB
Interference Assay of Spacers 37s and 94a
Ll.LtrB Insertion Assay in E. coli
Ll.LtrB Insertion Assay in P. putida
CRISPR/Cas9 Counterselection Assay in P. putida KT2440 and KT2440 ΔrecA
Calculation of Merged Targetrons/CRISPR/Cas9 Efficiency



Analysis of Ll.LtrB Insertion by Colony PCR
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acssynbio.1c00199.
List of oligonucleotides used in this study (Supplementary Table S1); list of plasmids used in this work (Supplementary Table S2); insertion frequencies of Ll.LtrB::Lux1 and Ll.LtrB::Lux4 in P. putida KT2440 WT and ΔrecA with no CRISPR/Cas9-mediated counterselection (Supplementary Table S3); insertion frequency of Ll.LtrB::LuxN intron in P. putida KT2440 WT with 5FOA CRISPR/Cas9-mediated counterselection (Supplementary Table S4); insertion frequency of Ll.LtrB::LuxN intron in P. putida KT2440 ΔrecA with 5FOA CRISPR/Cas9-mediated counterselection (Supplementary Table S5); construction and verification of intron delivery plasmids compatible with CRISPR/Cas9-mediated counterselection (Supplementary Methods); pSEVA plasmids for the expression of the Ll.LtrB intron in a wide range of Gram-negative bacteria (Supplementary Figure S1); preliminary approach to assess size-restriction of intron-mediated delivery using luxC fragments as a cargo and 5FOA counterselection in log-phase induced cells (Supplementary Figure S2); total intron insertion frequency without differentiating the cargo that was being delivered (Supplementary Figure S3); barcode generation with a PCR using 3′-overlapping 119-mer oligonucleotides (Supplementary Figure S4); application of targetrons as a barcode delivery system (Supplementary Figure S5); and design and test of Locus 1 and 2 spacers for CRISPR/Cas9-mediated counterselection of Ll.LtrB::B3 group II intron (Supplementary Figure S6) (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgments
The authors are indebted to Prof. Chris Miller (UC Denver) for critical reading of the manuscript. Figure 1 includes an image retrieved from https://biorender.com/ and used under subscription 89B77366-0004.
References
This article references 67 other publications.
- 1Nikel, P. I.; de Lorenzo, V. Pseudomonas putida as a functional chassis for industrial biocatalysis: From native biochemistry to trans-metabolism. Metab. Eng. 2018, 50, 142– 155, DOI: 10.1016/j.ymben.2018.05.005Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXpvVSlt7k%253D&md5=fac6e8937d2e4f3d0c26bcde01e6777cPseudomonas putida as a functional chassis for industrial biocatalysis: From native biochemistry to trans-metabolismNikel, Pablo I.; de Lorenzo, VictorMetabolic Engineering (2018), 50 (), 142-155CODEN: MEENFM; ISSN:1096-7176. (Elsevier B.V.)The itinerary followed by Pseudomonas putida from being a soil-dweller and plant colonizer bacterium to become a flexible and engineer-able platform for metabolic engineering stems from its natural lifestyle, which is adapted to harsh environmental conditions and all sorts of physicochem. stresses. Over the years, these properties have been capitalized biotechnol. owing to the expanding wealth of genetic tools designed for deep-editing the P. putida genome. A suite of dedicated vectors inspired in the core tenets of synthetic biol. have enabled to suppress many of the naturally-occurring undesirable traits native to this species while enhancing its many appealing properties, and also to import catalytic activities and attributes from other biol. systems. Much of the biotechnol. interest on P. putida stems from the distinct architecture of its central carbon metab. The native biochem. is naturally geared to generate reductive currency [i.e., NAD(P)H] that makes this bacterium a phenomenal host for redox-intensive reactions. In some cases, genetic editing of the indigenous biochem. network of P. putida (cis-metab.) has sufficed to obtain target compds. of industrial interest. Yet, the main value and promise of this species (in particular, strain KT2440) resides not only in its capacity to host heterologous pathways from other microorganisms, but also altogether artificial routes (trans-metab.) for making complex, new-to-Nature mols. A no. of examples are presented for substantiating the worth of P. putida as one of the favorite workhorses for sustainable manufg. of fine and bulk chems. in the current times of the 4th Industrial Revolution. The potential of P. putida to extend its rich native biochem. beyond existing boundaries is discussed and research bottlenecks to this end are also identified. These aspects include not just the innovative genetic design of new strains but also the incorporation of novel chem. elements into the extant biochem., as well as genomic stability and scaling-up issues.
- 2Nikel, P. I.; Martínez-García, E.; de Lorenzo, V. Biotechnological domestication of pseudomonads using synthetic biology. Nat. Rev. Microbiol. 2014, 12, 368– 379, DOI: 10.1038/nrmicro3253Google Scholar2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXmtlahtL8%253D&md5=d2cc342c2b17f5631572b664f5513acbBiotechnological domestication of pseudomonads using synthetic biologyNikel, Pablo I.; Martinez-Garcia, Esteban; de Lorenzo, VictorNature Reviews Microbiology (2014), 12 (5), 368-379CODEN: NRMACK; ISSN:1740-1526. (Nature Publishing Group)A review. Much of contemporary synthetic biol. research relies on the use of bacterial chassis for plugging-in and plugging-out genetic circuits and new-to-nature functionalities. However, the microorganisms that are the easiest to manipulate in the lab. are often suboptimal for downstream industrial applications, which can involve physicochem. stress and harsh operating conditions. In this Review, we advocate the use of environmental Pseudomonas strains as model organisms that are pre-endowed with the metabolic, physiol. and stress-endurance traits that are demanded by current and future synthetic biol. and biotechnol. needs.
- 3Kampers, L. F. C.; Volkers, R. J. M.; Martins Dos Santos, V. A. P. Pseudomonas putida KT2440 is HV1 certified, not GRAS. Microb. Biotechnol. 2019, 12, 845– 848, DOI: 10.1111/1751-7915.13443Google Scholar3https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BB3M3msleqsA%253D%253D&md5=01a803206ea94a4286be8ab8a039549cPseudomonas putida KT2440 is HV1 certified, not GRASKampers Linde F C; Volkers Rita J M; Martins Dos Santos Vitor A P; Martins Dos Santos Vitor A PMicrobial biotechnology (2019), 12 (5), 845-848 ISSN:.Pseudomonas putida is rapidly becoming a workhorse for industrial production due to its metabolic versatility, genetic accessibility and stress-resistance properties. The P. putida strain KT2440 is often described as Generally Regarded as Safe, or GRAS, indicating the strain is safe to use as food additive. This description is incorrect. P. putida KT2440 is classified by the FDA as HV1 certified, indicating it is safe to use in a P1 or ML1 environment.
- 4Kim, J.; Park, W. Oxidative stress response in Pseudomonas putida. Appl. Microbiol. Biotechnol. 2014, 98, 6933– 6946, DOI: 10.1007/s00253-014-5883-4Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhtVamtLbN&md5=de993728cab8409fdd59ed5c00fcb724Oxidative stress response in Pseudomonas putidaKim, Jisun; Park, WoojunApplied Microbiology and Biotechnology (2014), 98 (16), 6933-6946CODEN: AMBIDG; ISSN:0175-7598. (Springer)A review. Pseudomonas putida is widely distributed in nature and is capable of degrading various org. compds. due to its high metabolic versatility. The survival capacity of P. putida stems from its frequent exposure to various endogenous and exogenous oxidative stresses. Oxidative stress is an unavoidable consequence of interactions with various reactive oxygen species (ROS)-inducing agents existing in various niches. ROS could facilitate the evolution of bacteria by mutating genomes. Aerobic bacteria maintain defense mechanisms against oxidative stress throughout their evolution. To overcome the detrimental effects of oxidative stress, P. putida has developed defensive cellular systems involving induction of stress-sensing proteins and detoxification enzymes as well as regulation of oxidative stress response networks. Genetic responses to oxidative stress in P. putida differ markedly from those obsd. in Escherichia coli and Salmonella spp. Two major redox-sensing transcriptional regulators, SoxR and OxyR, are present and functional in the genome of P. putida. However, the novel regulators FinR and HexR control many genes belonging to the E. coli SoxR regulon. Oxidative stress can be generated by exposure to antibiotics, and iron homeostasis in P. putida is crucial for bacterial cell survival during treatment with antibiotics. This review highlights and summarizes current knowledge of oxidative stress in P. putida, as a model soil bacterium, together with recent studies from mol. genetics perspectives.
- 5Nikel, P. I.; Fuhrer, T.; Chavarría, M.; Sánchez-Pascuala, A.; Sauer, U.; de Lorenzo, V. Reconfiguration of metabolic fluxes in Pseudomonas putida as a response to sub-lethal oxidative stress. ISME J. 2021, 15, 1751– 1766, DOI: 10.1038/s41396-020-00884-9Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXitVSisLc%253D&md5=31d1780e11c16b4bbdbf09c008f4324dReconfiguration of metabolic fluxes in Pseudomonas putida as a response to sub-lethal oxidative stressNikel, Pablo I.; Fuhrer, Tobias; Chavarria, Max; Sanchez-Pascuala, Alberto; Sauer, Uwe; de Lorenzo, VictorISME Journal (2021), 15 (6), 1751-1766CODEN: IJSOCF; ISSN:1751-7362. (Nature Portfolio)As a frequent inhabitant of sites polluted with toxic chems., the soil bacterium and plant-root colonizer Pseudomonas putida can tolerate high levels of endogenous and exogenous oxidative stress. Yet, the ultimate reason of such phenotypic property remains largely unknown. To shed light on this question, metabolic network-wide routes for NADPH generation-the metabolic currency that fuels redox-stress quenching mechanisms-were inspected when P. putida KT2440 was challenged with a sub-lethal H2O2 dose as a proxy of oxidative conditions. 13C-tracer expts., metabolomics, and flux anal., together with the assessment of physiol. parameters and measurement of enzymic activities, revealed a substantial flux reconfiguration in oxidative environments. In particular, periplasmic glucose processing was rerouted to cytoplasmic oxidn., and the cyclic operation of the pentose phosphate pathway led to significant NADPH-forming fluxes, exceeding biosynthetic demands by ∼50%. The resulting NADPH surplus, in turn, fueled the glutathione system for H2O2 redn. These properties not only account for the tolerance of P. putida to environmental insults-some of which end up in the formation of reactive oxygen species-but they also highlight the value of this bacterial host as a platform for environmental bioremediation and metabolic engineering.
- 6Liao, J. C.; Mi, L.; Pontrelli, S.; Luo, S. Fuelling the future: Microbial engineering for the production of sustainable biofuels. Nat. Rev. Microbiol. 2016, 14, 288– 304, DOI: 10.1038/nrmicro.2016.32Google Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XltVaqsbk%253D&md5=fcd67bab889558ff6f918a5d70a9d6ecFuelling the future: microbial engineering for the production of sustainable biofuelsLiao, James C.; Mi, Luo; Pontrelli, Sammy; Luo, ShanshanNature Reviews Microbiology (2016), 14 (5), 288-304CODEN: NRMACK; ISSN:1740-1526. (Nature Publishing Group)Global climate change linked to the accumulation of greenhouse gases has caused concerns regarding the use of fossil fuels as the major energy source. To mitigate climate change while keeping energy supply sustainable, one soln. is to rely on the ability of microorganisms to use renewable resources for biofuel synthesis. In this Review, we discuss how microorganisms can be explored for the prodn. of next-generation biofuels, based on the ability of bacteria and fungi to use lignocellulose; through direct CO2 conversion by microalgae; using lithoautotrophs driven by solar electricity; or through the capacity of microorganisms to use methane generated from landfill. Furthermore, we discuss how to direct these substrates to the biosynthetic pathways of various fuel compds. and how to optimize biofuel prodn. by engineering fuel pathways and central metab.
- 7Jiménez, J. I.; Miñambres, B.; García, J. L.; Díaz, E. Genomic analysis of the aromatic catabolic pathways from Pseudomonas putida KT2440. Environ. Microbiol. 2002, 4, 824– 841, DOI: 10.1046/j.1462-2920.2002.00370.xGoogle Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXkvVGrtg%253D%253D&md5=1922f12af80e022107400b58fa47f419Genomic analysis of the aromatic catabolic pathways from Pseudomonas putida KT2440Jimenez, Jose Ignacio; Minambres, Baltasar; Garcia, Jose Luis; Diaz, EduardoEnvironmental Microbiology (2002), 4 (12), 824-841CODEN: ENMIFM; ISSN:1462-2912. (Blackwell Science Ltd.)Anal. of the catabolic potential of Pseudomonas putida KT2440 against a wide range of natural arom. compds. and sequence comparisons with the entire genome of this microorganism predicted the existence of at least four main pathways for the catabolism of central arom. intermediates, i.e., the protocatechuate (pca genes) and catechol (cat genes) branches of the β-ketoadipate pathway, the homogentisate pathway (hmg/fah/mai genes) and the phenylacetate pathway (pha genes). Two addnl. gene clusters that might be involved in the catabolism of N-heterocyclic arom. compds. (nic cluster) and in a central meta-cleavage pathway (pcm genes) were also identified. Furthermore, the genes encoding the peripheral pathways for the catabolism of p-hydroxybenzoate (pob), benzoate (ben), quinate (qui), phenylpropenoid compds. (fcs, ech, vdh, cal, van, acd and acs), phenylalanine and tyrosine (phh, hpd) and n-phenylalkanoic acids (fad) were mapped in the chromosome of P. putida KT2440. Although a repetitive extragenic palindromic (REP) element is usually assocd. with the gene clusters, a supraoperonic clustering of catabolic genes that channel different arom. compds. into a common central pathway (catabolic island) was not obsd. in P. putida KT2440. The global view on the mineralization of arom. compds. by P. putida KT2440 will facilitate the rational manipulation of this strain for improving biodegrdn./biotransformation processes, and reveals this bacterium as a useful model system for studying biochem., genetic, evolutionary and ecol. aspects of the catabolism of arom. compds.
- 8Ramos, J. L.; Duque, E.; Huertas, M. J.; Haidour, A. Isolation and expansion of the catabolic potential of a Pseudomonas putida strain able to grow in the presence of high concentrations of aromatic hydrocarbons. J. Bacteriol. 1995, 177, 3911– 3916, DOI: 10.1128/jb.177.14.3911-3916.1995Google Scholar8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXmvV2gsb0%253D&md5=92a84eea5b4f2a8abba19a91b534c80bIsolation and expansion of the catabolic potential of a Pseudomonas putida strain able to grow in the presence of high concentrations of aromatic hydrocarbonsRamos, Juan L.; Duque, Estrella; Huertas, Maria-Jose; Haidour, AliJournal of Bacteriology (1995), 177 (14), 3911-16CODEN: JOBAAY; ISSN:0021-9193. (American Society for Microbiology)Pseudomonas putida DOT-T1 was isolated after enrichment on minimal medium with 1% (vol/vol) toluene as the sole C source. The strain was able to grow in the presence of 90% (vol/vol) toluene and was tolerant to org. solvents whose log Pow (octanol/water partition coeff.) was higher than 2.3. Solvent tolerance was inducible, as bacteria grown in the absence of toluene required an adaptation period before growth restarted. Mg2+ ions in the culture medium improved solvent tolerance. Electron micrographs showed that cells growing on high concns. of toluene exhibited a wider periplasmic space than cells growing in the absence of toluene and preserved the outer membrane integrity. Polarog. studies and the accumulation of pathway intermediates showed that the strain used the toluene-4-monooxygenase pathway to catabolize toluene. Although the strain also thrived in high concns. of m- and p-xylene, these hydrocarbons could not be used as the sole C source for growth. The catabolic potential of the isolate was expanded to include m- and p-xylene and related hydrocarbons by transfer of the TOL plasmid pWW0-Km.
- 9Martínez, I.; Mohamed, M. E. S.; Rozas, D.; García, J. L.; Díaz, E. Engineering synthetic bacterial consortia for enhanced desulfurization and revalorization of oil sulfur compounds. Metab. Eng. 2016, 35, 46– 54, DOI: 10.1016/j.ymben.2016.01.005Google Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28Xhtlalu70%253D&md5=6021bcdce6cb854cabf99a30150e080aEngineering synthetic bacterial consortia for enhanced desulfurization and revalorization of oil sulfur compoundsMartinez, Igor; Mohamed, Magdy El-Said; Rozas, Daniel; Garcia, Jose Luis; Diaz, EduardoMetabolic Engineering (2016), 35 (), 46-54CODEN: MEENFM; ISSN:1096-7176. (Elsevier B. V.)The 4S pathway is the most studied bioprocess for the removal of the recalcitrant sulfur of arom. heterocycles present in fuels. It consists of three sequential functional units, encoded by the dszABCD genes, through which the model compd. dibenzothiophene (DBT) is transformed into the sulfur-free 2-hydroxybiphenyl (2HBP) mol. In this work, a set of synthetic dsz cassettes were implanted in Pseudomonas putida KT2440, a model bacterial "chassis" for metabolic engineering studies. The complete dszB1A1C1-D1 cassette behaved as an attractive alternative - to the previously constructed recombinant dsz cassettes - for the conversion of DBT into 2HBP. Refactoring the 4S pathway by the use of synthetic dsz modules encoding individual 4S pathway reactions revealed unanticipated traits, e.g., the 4S intermediate 2HBP-sulfinate (HBPS) behaves as an inhibitor of the Dsz monooxygenases, and once secreted from the cells it cannot be further taken up. That issue should be addressed for the rational design of more efficient biocatalysts for DBT bioconversions. In this sense, the construction of synthetic bacterial consortia to compartmentalize the 4S pathway into different cell factories for individual optimization was shown to enhance the conversion of DBT into 2HBP, overcome the inhibition of the Dsz enzymes by the 4S intermediates, and enable efficient prodn. of unattainable high added value intermediates, e.g., HBPS, that are difficult to obtain using the current monocultures.
- 10Martínez-García, E.; Nikel, P. I.; Aparicio, T.; de Lorenzo, V. Pseudomonas 2.0: Genetic upgrading of P. putida KT2440 as an enhanced host for heterologous gene expression. Microb. Cell Fact. 2014, 13, 159 DOI: 10.1186/s12934-014-0159-3Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXit1Oqt7g%253D&md5=dfebf084cef553ec5d6526f9fbbd504fPseudomonas 2.0: genetic upgrading of P. putida KT2440 as an enhanced host for heterologous gene expressionMartinez-Garcia, Esteban; Nikel, Pablo I.; Aparicio, Tomas; de Lorenzo, VictorMicrobial Cell Factories (2014), 13 (), 159/1-159/33CODEN: MCFICT; ISSN:1475-2859. (BioMed Central Ltd.)Background: Because of its adaptability to sites polluted with toxic chems., the model soil bacterium Pseudomonas putida is naturally endowed with a no. of metabolic and stress-endurance qualities which have considerable value for hosting energy-demanding and redox reactions thereof. The growing body of knowledge on P. putida strain KT2440 has been exploited for the rational design of a deriv. strain in which the genome has been heavily edited in order to construct a robust microbial cell factory. Results: Eleven non-adjacent genomic deletions, which span 300 genes (i.e., 4.3% of the entire P. putida KT2440 genome), were eliminated; thereby enhancing desirable traits and eliminating attributes which are detrimental in an expression host. Since ATP and NAD(P)H availability - as well as genetic instability, are generally considered to be major bottlenecks for the performance of platform strains, a suite of functions that drain high-energy phosphate from the cells and/or consume NAD(P)H were targeted in particular, the whole flagellar machinery. Four prophages, two transposons, and three components of DNA restriction modification systems were eliminated as well. The resulting strain (P. putida EM383) displayed growth properties (i.e., lag times, biomass yield, and specific growth rates) clearly superior to the precursor wild-type strain KT2440. Furthermore, it tolerated endogenous oxidative stress, acquired and replicated exogenous DNA, and survived better in stationary phase. The performance of a bi-cistronic GFP-LuxCDABE reporter system as a proxy of combined metabolic vitality, revealed that the deletions in P. putida strain EM383 brought about an increase of >50% in the overall physiol. vigour. Conclusion: The rationally modified P. putida strain allowed for the better functional expression of implanted genes by directly improving the metabolic currency that sustains the gene expression flow, instead of resorting to the classical genetic approaches (e.g., increasing the promoter strength in the DNA constructs of interest).
- 11Lieder, S.; Nikel, P. I.; de Lorenzo, V.; Takors, R. Genome reduction boosts heterologous gene expression in Pseudomonas putida. Microb. Cell Fact. 2015, 14, 23 DOI: 10.1186/s12934-015-0207-7Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC2MjlvVeqtA%253D%253D&md5=1b3829a02ebcccfb6bd3f5dfc0299255Genome reduction boosts heterologous gene expression in Pseudomonas putidaLieder Sarah; Takors Ralf; Nikel Pablo I; de Lorenzo VictorMicrobial cell factories (2015), 14 (), 23 ISSN:.BACKGROUND: The implementation of novel platform organisms to be used as microbial cell factories in industrial applications is currently the subject of intense research. Ongoing efforts include the adoption of Pseudomonas putida KT2440 variants with a reduced genome as the functional chassis for biotechnological purposes. In these strains, dispensable functions removed include flagellar motility (1.1% of the genome) and a number of open reading frames expected to improve genotypic and phenotypic stability of the cells upon deletion (3.2% of the genome). RESULTS: In this study, two previously constructed multiple-deletion P. putida strains were systematically evaluated as microbial cell factories for heterologous protein production and compared to the parental bacterium (strain KT2440) with regards to several industrially-relevant physiological traits. Energetic parameters were quantified at different controlled growth rates in continuous cultivations and both strains had a higher adenosine triphosphate content, increased adenylate energy charges, and diminished maintenance demands than the wild-type strain. Under all the conditions tested the mutants also grew faster, had enhanced biomass yields and showed higher viability, and displayed increased plasmid stability than the parental strain. In addition to small-scale shaken-flask cultivations, the performance of the genome-streamlined strains was evaluated in larger scale bioreactor batch cultivations taking a step towards industrial growth conditions. When the production of the green fluorescent protein (used as a model heterologous protein) was assessed in these cultures, the mutants reached a recombinant protein yield with respect to biomass up to 40% higher than that of P. putida KT2440. CONCLUSIONS: The two streamlined-genome derivatives of P. putida KT2440 outcompeted the parental strain in every industrially-relevant trait assessed, particularly under the working conditions of a bioreactor. Our results demonstrate that these genome-streamlined bacteria are not only robust microbial cell factories on their own, but also a promising foundation for further biotechnological applications.
- 12Silva-Rocha, R.; Martínez-García, E.; Calles, B.; Chavarría, M.; Arce-Rodríguez, A.; De Las Heras, A.; Páez-Espino, A. D.; Durante-Rodríguez, G.; Kim, J.; Nikel, P. I.; Platero, R.; De Lorenzo, V. The Standard European Vector Architecture (SEVA): A coherent platform for the analysis and deployment of complex prokaryotic phenotypes. Nucleic Acids Res. 2013, 41, D666– D675, DOI: 10.1093/nar/gks1119Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhvV2ktrzF&md5=3ddd60ae5099d577aa6b688813a5605bThe Standard European Vector Architecture (SEVA): a coherent platform for the analysis and deployment of complex prokaryotic phenotypesSilva-Rocha, Rafael; Martinez-Garcia, Esteban; Calles, Belen; Chavarria, Max; Arce-Rodriguez, Alejandro; de las Heras, Aitor; Paez-Espino, A. David; Durante-Rodriguez, Gonzalo; Kim, Juhyun; Nikel, Pablo I.; Platero, Raul; de Lorenzo, VictorNucleic Acids Research (2013), 41 (D1), D666-D675CODEN: NARHAD; ISSN:0305-1048. (Oxford University Press)The Std. European Vector Architecture' database (SEVA-DB, http://seva.cnb.csic.es) was conceived as a user-friendly, web-based resource and a material clone repository to assist in the choice of optimal plasmid vectors for de-constructing and re-constructing complex prokaryotic phenotypes. The SEVA-DB adopts simple design concepts that facilitate the swapping of functional modules and the extension of genome engineering options to microorganisms beyond typical lab. strains. Under the SEVA std., every DNA portion of the plasmid vectors is minimized, edited for flaws in their sequence and/or functionality, and endowed with phys. connectivity through three inter-segment insulators that are flanked by fixed, rare restriction sites. Such a scaffold enables the exchangeability of multiple origins of replication and diverse antibiotic selection markers to shape a frame for their further combination with a large variety of cargo modules that can be used for varied end-applications. The core collection of constructs that are available at the SEVA-DB has been produced as a starting point for the further expansion of the formatted vector platform. We argue that adoption of the SEVA format can become a shortcut to fill the phenomenal gap between the existing power of DNA synthesis and the actual engineering of predictable and efficacious bacteria.
- 13Martínez-García, E.; de Lorenzo, V. Engineering multiple genomic deletions in Gram-negative bacteria: Analysis of the multi-resistant antibiotic profile of Pseudomonas putida KT2440. Environ. Microbiol. 2011, 13, 2702– 2716, DOI: 10.1111/j.1462-2920.2011.02538.xGoogle Scholar13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhsVKhtbfP&md5=8a2204c9c029721125df30a7c2ac80e2Engineering multiple genomic deletions in gram-negative bacteria: analysis of the multi-resistant antibiotic profile of Pseudomonas putida KT2440Martinez-Garcia, Esteban; de Lorenzo, VictorEnvironmental Microbiology (2011), 13 (10), 2702-2716CODEN: ENMIFM; ISSN:1462-2912. (Wiley-Blackwell)The genome of the soil bacterium Pseudomonas putida strain KT2440 has been erased of various determinants of resistance to antibiotics encoded in its extant chromosome. To this end, the authors employed a coherent genetic platform that allowed the precise deletion of multiple genomic segments in a large variety of Gram-neg. bacteria including (but not limited to) P. putida. The method is based on the obligatory recombination between free-ended homologous DNA sequences that are released as linear fragments generated upon the cleavage of the chromosome with unique I-SceI sites, added to the segment of interest by the vector system. Despite the potential for a SOS response brought about by the appearance of double stranded DNA breaks during the process, fluctuation expts. revealed that the procedure did not increase mutation rates - perhaps due to the protection exerted by I-SceI bound to the otherwise naked DNA termini. With this tool in hand the authors made sequential deletions of genes mexC, mexE, ttgA and ampC in the genome of the target bacterium, orthologues of which are known to det. various degrees of antibiotic resistance in diverse microorganisms. Inspection of the corresponding phenotypes demonstrated that the efflux pump encoded by ttgA sufficed to endow P. putida with a high-level of tolerance to β-lactams, chloramphenicol and quinolones, but had little effect on, e.g. aminoglycosides. Anal. of the mutants revealed also a considerable diversity in the manifestation of the resistance phenotype within the population and suggested a degree of synergism between different pumps. The directed edition of the P. putida chromosome shown here not only enhances the amenability of this bacterium to deep genomic engineering, but also validates the corresponding approach for similar handlings of a large variety of Gram-neg. microorganisms.
- 14Martínez-García, E.; De Lorenzo, V. Transposon-based and plasmid-based genetic tools for editing genomes of Gram-negative bacteria. Methods Mol. Biol. 2012, 813, 267– 283Google Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhs1Sqs7jK&md5=d8dc7fef841f41065b1e0c236b3339d8Transposon-based and plasmid-based genetic tools for editing genomes of gram-negative bacteriaMartinez-Garcia, Esteban; de Lorenzo, VictorMethods in Molecular Biology (New York, NY, United States) (2012), 813 (Synthetic Gene Networks), 267-283CODEN: MMBIED; ISSN:1064-3745. (Springer)A good part of the contemporary synthetic biol. agenda aims at reprogramming microorganisms to enhance existing functions and/or perform new tasks. Moreover, the functioning of complex regulatory networks, or even a single gene, is revealed only when perturbations are entered in the corresponding dynamic systems and the outcome monitored. These endeavors rely on the availability of genetic tools to successfully modify a la carte the chromosome of target bacteria. Key aspects to this end include the removal of undesired genomic segments, systems for the prodn. of directed mutants and allelic replacements, random mutant libraries to discover new functions, and means to stably implant larger genetic networks into the genome of specific hosts. The list of gram-neg. species that are appealing for such genetic refactoring operations is growingly expanding. However, the repertoire of available mol. techniques to do so is very limited beyond Escherichia coli. In this chapter, utilization of novel tools is described (exemplified in two plasmids systems: pBAM1 and pEMG) tailored for facilitating chromosomal engineering procedures in a wide variety of gram-neg. microorganisms.
- 15Martínez-García, E.; Aparicio, T.; de Lorenzo, V.; Nikel, P. I. New transposon tools tailored for metabolic engineering of Gram-negative microbial cell factories. Front. Bioeng. Biotechnol. 2014, 2, 46 DOI: 10.3389/fbioe.2014.00046Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC2M3nsVOrtw%253D%253D&md5=e535341c401daaac2882e3d09ba452b3New transposon tools tailored for metabolic engineering of gram-negative microbial cell factoriesMartinez-Garcia Esteban; Aparicio Tomas; de Lorenzo Victor; Nikel Pablo IFrontiers in bioengineering and biotechnology (2014), 2 (), 46 ISSN:2296-4185.Re-programming microorganisms to modify their existing functions and/or to bestow bacteria with entirely new-to-Nature tasks have largely relied so far on specialized molecular biology tools. Such endeavors are not only relevant in the burgeoning metabolic engineering arena but also instrumental to explore the functioning of complex regulatory networks from a fundamental point of view. A la carte modification of bacterial genomes thus calls for novel tools to make genetic manipulations easier. We propose the use of a series of new broad-host-range mini-Tn5-vectors, termed pBAMDs, for the delivery of gene(s) into the chromosome of Gram-negative bacteria and for generating saturated mutagenesis libraries in gene function studies. These delivery vectors endow the user with the possibility of easy cloning and subsequent insertion of functional cargoes with three different antibiotic-resistance markers (kanamycin, streptomycin, and gentamicin). After validating the pBAMD vectors in the environmental bacterium Pseudomonas putida KT2440, their use was also illustrated by inserting the entire poly(3-hydroxybutyrate) (PHB) synthesis pathway from Cupriavidus necator in the chromosome of a phosphotransacetylase mutant of Escherichia coli. PHB is a completely biodegradable polyester with a number of industrial applications that make it attractive as a potential replacement of oil-based plastics. The non-selective nature of chromosomal insertions of the biosynthetic genes was evidenced by a large landscape of PHB synthesis levels in independent clones. One clone was selected and further characterized as a microbial cell factory for PHB accumulation, and it achieved polymer accumulation levels comparable to those of a plasmid-bearing recombinant. Taken together, our results demonstrate that the new mini-Tn5-vectors can be used to confer interesting phenotypes in Gram-negative bacteria that would be very difficult to engineer through direct manipulation of the structural genes.
- 16Aparicio, T.; Nyerges, A.; Martínez-García, E.; de Lorenzo, V. High-Efficiency Multi-site Genomic Editing of Pseudomonas putida through Thermoinducible ssDNA Recombineering. iScience 2020, 23, 100946 DOI: 10.1016/j.isci.2020.100946Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXmsl2mtLg%253D&md5=7b4bae6ce59f76035cfa133e40ac0be5High-Efficiency Multi-site Genomic Editing of Pseudomonas putida through Thermoinducible ssDNA RecombineeringAparicio, Tomas; Nyerges, Akos; Martinez-Garcia, Esteban; de Lorenzo, VictoriScience (2020), 23 (3), 100946CODEN: ISCICE; ISSN:2589-0042. (Elsevier B.V.)Application of single-stranded DNA recombineering for genome editing of species other than enterobacteria is limited by the efficiency of the recombinase and the action of endogenous mismatch repair (MMR) systems. In this work we have set up a genetic system for entering multiple changes in the chromosome of the biotechnol. relevant strain EM42 of Pseudomononas putida. To this end high-level heat-inducible co-transcription of the rec2 recombinase and P. putida's allele mutLPPE36K was designed under the control of the PL/cI857 system. Cycles of short thermal shifts followed by transformation with a suite of mutagenic oligos delivered different types of genomic changes at frequencies up to 10% per single modification. The same approach was instrumental to super-diversify short chromosomal portions for creating libraries of functional genomic segments-e.g., ribosomal-binding sites. These results enabled multiplexing of genome engineering of P. putida, as required for metabolic reprogramming of this important synthetic biol. chassis.
- 17Martin-Pascual, M.; Batianis, C.; Bruinsma, L.; Asin-Garcia, E.; Garcia-Morales, L.; Weusthuis, R. A.; van Kranenburg, R.; Martins Dos Santos, V. A. P. A navigation guide of synthetic biology tools for Pseudomonas putida. Biotechnol. Adv. 2021, 49, 107732 DOI: 10.1016/j.biotechadv.2021.107732Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhs1yrt7bL&md5=3012891782213362ce45205f20b9ddfeA navigation guide of synthetic biology tools for Pseudomonas putidaMartin-Pascual, Maria; Batianis, Christos; Bruinsma, Lyon; Asin-Garcia, Enrique; Garcia-Morales, Luis; Weusthuis, Ruud A.; van Kranenburg, Richard; Martins dos Santos, Vitor A. P.Biotechnology Advances (2021), 49 (), 107732CODEN: BIADDD; ISSN:0734-9750. (Elsevier Inc.)Pseudomonas putida is a microbial chassis of huge potential for industrial and environmental biotechnol., owing to its remarkable metabolic versatility and ability to sustain difficult redox reactions and operational stresses, among other attractive characteristics. A wealth of genetic and in silico tools have been developed to enable the unravelling of its physiol. and improvement of its performance. However, the rise of this microbe as a promising platform for biotechnol. applications has resulted in diversification of tools and methods rather than standardization and convergence. As a consequence, multiple tools for the same purpose have been generated, while most of them have not been embraced by the scientific community, which has led to compartmentalization and inefficient use of resources. Inspired by this and by the substantial increase in popularity of P. putida, we aim herein to bring together and assess all currently available (wet and dry) synthetic biol. tools specific for this microbe, focusing on the last 5 years. We provide information on the principles, functionality, advantages and limitations, with special focus on their use in metabolic engineering. Addnl., we compare the tool portfolio for P. putida with those for other bacterial chassis and discuss potential future directions for tool development. Therefore, this review is intended as a ref. guide for experts and new 'users' of this promising chassis.
- 18Mörl, M.; Niemer, I.; Schmelzer, C. New reactions catalyzed by a group II intron ribozyme with RNA and DNA substrates. Cell 1992, 70, 803– 810, DOI: 10.1016/0092-8674(92)90313-2Google Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADyaK38zos1Gquw%253D%253D&md5=0de88e6dcdb98d1f7c6c68cdac24d25eNew reactions catalyzed by a group II intron ribozyme with RNA and DNA substratesMorl M; Niemer I; Schmelzer CCell (1992), 70 (5), 803-10 ISSN:0092-8674.Here we describe three novel reactions of the self-splicing group II intron bI1 (the first intron of the COB gene of yeast mitochondria) demonstrating its catalytic versatility: reversal of the first step of the self-splicing reaction catalyzed by a linear form of the intron utilizing the energy of a phosphoanhydride bond for transesterification, ligation of a single-stranded DNA to an RNA, and cleavage of a single-stranded DNA substrate. These results have the following evolutionary implications: use of the alpha-beta bond of a terminal triphosphate for transesterification suggests that an RNA RNA replicase could use mononucleotide triphosphates as precursors, and cleavage of single-stranded DNA and DNA-RNA ligation suggests that excised group II introns might integrate directly into DNA without prior reverse transcription.
- 19Lazowska, J.; Meunier, B.; Macadre, C. Homing of a group II intron in yeast mitochondrial DNA is accompanied by unidirectional co-conversion of upstream-located markers. EMBO J. 1994, 13, 4963– 4972, DOI: 10.1002/j.1460-2075.1994.tb06823.xGoogle Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXhvF2hsrY%253D&md5=c7313b7d97832fa83b5ed5c1ba63683bHoming of a group II intron in yeast mitochondrial DNA is accompanied by unidirectional co-conversion of upstream-located markersLazowska, Jaga; Meunier, Brigitte; Macadre, CatherineEMBO Journal (1994), 13 (20), 4963-72CODEN: EMJODG; ISSN:0261-4189. (Oxford University Press)Group II introns ai1 and ai2 of the Saccharomyces cerevisiae mitochondrial COX1 gene encode proteins having a dual function (maturase and reverse transcriptase) and are mobile genetic elements. By construction of adequate donor genomes, we demonstrate that each of them is self-sufficient and practises homing in the absence of homing-type endonucleases encoded by either group I introns or the ENS2 gene. Each of the S. cerevisiae group II self-mobile introns was tested for its ability to invade mitochondrial DNA (mtDNA) from two related Saccharomyces species. Surprisingly, only ai2 was obsd. to integrate into both genomes. The non-mobility of ai1 was clearly correlated with some polymorphic changes occurring in sequences flanking its insertion sites in the recipient mtDNAs. Importantly, studies of the behavior of these introns in interspecific crosses demonstrate that flanking marker co-conversion accompanying group II intron homing is unidirectional and efficient only in the 3' to 5' direction towards the upstream exon. Thus, the polar co-conversion and dependence of the splicing proficiency of the intron reported previously by us are hallmarks of group II intron homing, which significantly distinguish it from the strictly DNA-based group I intron homing and strictly RNA-based group II intron transposition.
- 20Saldanha, R.; Chen, B.; Wank, H.; Matsuura, M.; Edwards, J.; Lambowitz, A. M. RNA and protein catalysis in group II intron splicing and mobility reactions using purified components. Biochemistry 1999, 38, 9069– 9083, DOI: 10.1021/bi982799lGoogle Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXjvVCqtro%253D&md5=2136004d61355a78d3ad72c72f646545RNA and Protein Catalysis in Group II Intron Splicing and Mobility Reactions Using Purified ComponentsSaldanha, Roland; Chen, Bing; Wank, Herbert; Matsuura, Manabu; Edwards, Judy; Lambowitz, Alan M.Biochemistry (1999), 38 (28), 9069-9083CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)Group II introns encode proteins with reverse transcriptase activity. These proteins also promote RNA splicing (maturase activity) and then, with the excised intron, form a site-specific DNA endonuclease that promotes intron mobility by reverse splicing into DNA followed by target DNA-primed reverse transcription. Here, we used an Escherichia coli expression system for the Lactococcus lactis group II intron Ll.LtrB to show that the intron-encoded protein (LtrA) alone is sufficient for maturase activity, and that RNP particles contg. only the LtrA protein and excised intron RNA have site-specific DNA endonuclease and target DNA-primed reverse transcriptase activity. Detailed anal. of the splicing reaction indicates that LtrA is an intron-specific splicing factor that binds to unspliced precursor RNA with a Kd of ≤0.12 pM at 30°. This binding occurs in a rapid bimol. reaction, which is followed by a slower step, presumably an RNA conformational change, required for splicing to occur. Our results constitute the first biochem. anal. of protein-dependent splicing of a group II intron and demonstrate that a single intron-encoded protein can interact with the intron RNA to carry out a coordinated series of reactions leading to splicing and mobility.
- 21Matsuura, M.; Noah, J. W.; Lambowitz, A. M. Mechanism of maturase-promoted group II intron splicing. EMBO J. 2001, 20, 7259– 7270, DOI: 10.1093/emboj/20.24.7259Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XpsF2h&md5=18a1b077ef7c4155b12d2fc0bef65540Mechanism of maturase-promoted group II intron splicingMatsuura, Manabu; Noah, James W.; Lambowitz, Alan M.EMBO Journal (2001), 20 (24), 7259-7270CODEN: EMJODG; ISSN:0261-4189. (Oxford University Press)Mobile group II introns encode reverse transcriptases that also function as intron-specific splicing factors (maturases). We showed previously that the reverse transcriptase/maturase encoded by the Lactococcus lactis Ll.LtrB intron has a high affinity binding site at the beginning of its own coding region in an idiosyncratic structure, DIVa. Here, we identify potential secondary binding sites in conserved regions of the catalytic core and show via chem. modification expts. that binding of the maturase induces the formation of key tertiary interactions required for RNA splicing. The interaction with conserved as well as idiosyneratic regions explains how maturases in some organisms could evolve into general group II intron splicing factors, potentially mirroring a key step in the evolution of spliceosomal introns.
- 22Lambowitz, A. M.; Zimmerly, S. Group II introns: Mobile ribozymes that invade DNA. Cold Spring Harbor Perspect. Biol. 2011, 3, a003616 DOI: 10.1101/cshperspect.a003616Google ScholarThere is no corresponding record for this reference.
- 23Perutka, J.; Wang, W.; Goerlitz, D.; Lambowitz, A. M. Use of Computer-designed Group II Introns to Disrupt Escherichia coli DExH/D-box Protein and DNA Helicase Genes. J. Mol. Biol. 2004, 336, 421– 439, DOI: 10.1016/j.jmb.2003.12.009Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXntVKrsw%253D%253D&md5=d7cde4aede8256f94d4df7eeed98ed2bUse of Computer-designed Group II Introns to Disrupt Escherichia coli DExH/D-box Protein and DNA Helicase GenesPerutka, Jiri; Wang, Wenjun; Goerlitz, David; Lambowitz, Alan M.Journal of Molecular Biology (2004), 336 (2), 421-439CODEN: JMOBAK; ISSN:0022-2836. (Elsevier)Mobile group II introns are site-specific retroelements that use a novel mobility mechanism in which the excised intron RNA inserts directly into a DNA target site and is then reverse transcribed by the assocd. intron-encoded protein. Because the DNA target site is recognized primarily by base-pairing of the intron RNA with only a small no. of positions recognized by the protein, it has been possible to develop group II introns into a new type of gene targeting vector ("targetron"), which can be reprogrammed to insert into desired DNA targets simply by modifying the intron RNA. Here, we used databases of retargeted Lactococcus lactis Ll.LtrB group II introns and a compilation of nucleotide frequencies at active target sites to develop an algorithm that predicts optimal Ll.LtrB intron-insertion sites and designs primers for modifying the intron to insert into those sites. In a test of the algorithm, we designed one or two targetrons to disrupt each of 28 Escherichia coli genes encoding DExH/D-box and DNA helicase-related proteins and tested for the desired disruptants by PCR screening of 100 colonies. In 21 cases, we obtained disruptions at frequencies of 1-80% without selection, and in six other cases, where disruptants were not identified in the initial PCR screen, we readily obtained specific disruptions by using the same targetrons with a retrotransposition-activated selectable marker. Only one DExH/D-box protein gene, secA, which was known to be essential, did not give viable disruptants. The apparent dispensability of DExH/D-box proteins in E. coli contrasts with the situation in yeast, where the majority of such proteins are essential. The methods developed here should permit the rapid and efficient disruption of any bacterial gene, the computational anal. provides new insight into group II intron target site recognition, and the set of E. coli DExH/D-box protein and DNA helicase disruptants should be useful for analyzing the function of these proteins.
- 24Heap, J. T.; Pennington, O. J.; Cartman, S. T.; Carter, G. P.; Minton, N. P. The ClosTron: A universal gene knock-out system for the genus Clostridium. J. Microbiol. Methods 2007, 70, 452– 464, DOI: 10.1016/j.mimet.2007.05.021Google Scholar24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXpsFKjsLk%253D&md5=5ad07920edc8de3b8f622522223016aaThe ClosTron: A universal gene knock-out system for the genus ClostridiumHeap, John T.; Pennington, Oliver J.; Cartman, Stephen T.; Carter, Glen P.; Minton, Nigel P.Journal of Microbiological Methods (2007), 70 (3), 452-464CODEN: JMIMDQ; ISSN:0167-7012. (Elsevier B.V.)Progress in exploiting clostridial genome information has been severely impeded by a general lack of effective methods for the directed inactivation of specific genes. Those few mutants that have been generated have been almost exclusively derived by single crossover integration of a replication-deficient or defective plasmid by homologous recombination. The mutants created are therefore unstable. Here we have adapted a mutagenesis system based on the mobile group II intron from the ltrB gene of Lactococcus lactis (Ll.ltrB) to function in clostridial hosts. Integrants are readily selected on the basis of acquisition of resistance to erythromycin, and are generated from start to finish in as little as 10 to 14 days. Unlike single crossover plasmid integrants, the mutants are extremely stable. The system has been used to make 6 mutants of Clostridium acetobutylicum and 5 of Clostridium difficile, exceeding the no. of published mutants ever generated in these species. Genes have also been inactivated for the first time in Clostridium botulinum and Clostridium sporogenes, suggesting the system will be universally applicable to the genus. The procedure is highly efficient and reproducible, and should revolutionize functional genomic studies in clostridia.
- 25Akhtar, P.; Khan, S. A. Two independent replicons can support replication of the anthrax toxin-encoding plasmid pXO1 of Bacillus anthracis. Plasmid 2012, 67, 111– 117, DOI: 10.1016/j.plasmid.2011.12.012Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xks12jtrY%253D&md5=e29997ac16445fd79197fc6dbd55ccb6Two independent replicons can support replication of the anthrax toxin-encoding plasmid pXO1 of Bacillus anthracisAkhtar, Parvez; Khan, Saleem A.Plasmid (2012), 67 (2), 111-117CODEN: PLSMDX; ISSN:0147-619X. (Elsevier)The large pXO1 plasmid (181.6 kb) of Bacillus anthracis encodes the anthrax toxin proteins. Previous studies have shown that two sep. regions of pXO1 can support replication of pXO1 miniplasmids when introduced into plasmid-less strains of this organism. No information is currently available on the ability of the above two replicons, termed RepX and ORFs 14/16 replicons, to support replication of the full-length pXO1 plasmid. We generated mutants of the full-length pXO1 plasmid in which either the RepX or the ORFs 14/16 replicon was inactivated by TargeTron insertional mutagenesis. Plasmid pXO1 derivs. contg. only the RepX or the ORFs 14/16 replicon were able to replicate when introduced into a plasmid-less B. anthracis strain. Plasmid copy no. anal. showed that the ORFs 14/16 replicon is more efficient than the RepX replicon. Our studies demonstrate that both the RepX and ORFs 14/16 replicons can independently support the replication of the full-length pXO1 plasmid.
- 26Frazier, C. L.; San Filippo, J.; Lambowitz, A. M.; Mills, D. A. Genetic manipulation of Lactococcus lactis by using targeted group II introns: Generation of stable insertions without selection. Appl. Environ. Microbiol. 2003, 69, 1121– 1128, DOI: 10.1128/AEM.69.2.1121-1128.2003Google Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXhtF2it7w%253D&md5=70f040e4515aded320ec667e68c53cceGenetic manipulation of Lactococcus lactis by using targeted group II introns: Generation of stable insertions without selectionFrazier, Courtney L.; San Filippo, Joseph; Lambowitz, Alan M.; Mills, David A.Applied and Environmental Microbiology (2003), 69 (2), 1121-1128CODEN: AEMIDF; ISSN:0099-2240. (American Society for Microbiology)Despite their com. importance, there are relatively few facile methods for genomic manipulation of the lactic acid bacteria. Here, the lactococcal group II intron, Ll.ltrB, was targeted to insert efficiently into genes encoding malate decarboxylase (mleS) and tetracycline resistance (tetM) within the Lactococcus lactis genome. Integrants were readily identified and maintained in the absence of a selectable marker. Since splicing of the Ll.ltrB intron depends on the intron-encoded protein, targeted invasion with an intron lacking the intron open reading frame disrupted TetM and MleS function, and MleS activity could be partially restored by expressing the intron-encoded protein in trans. Restoration of splicing from intron variants lacking the intron-encoded protein illustrates how targeted group II introns could be used for conditional expression of any gene. Furthermore, the modified Ll.ltrB intron was used to sep. deliver a phage resistance gene (abiD) and a tetracycline resistance marker (tetM) into mleS, without the need for selection to drive the integration or to maintain the integrant. Our findings demonstrate the utility of targeted group II introns as a potential food-grade mechanism for delivery of industrially important traits into the genomes of lactococci.
- 27Plante, I.; Cousineau, B. Restriction for gene insertion within the Lactococcus lactis Ll.LtrB group II intron. RNA 2006, 12, 1980– 1992, DOI: 10.1261/rna.193306Google Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XhtFygsL%252FF&md5=12308f25b4b51b486da6ce6398af0b73Restriction for gene insertion within the Lactococcus lactis Ll.LtrB group II intronPlante, Isabelle; Cousineau, BenoitRNA (2006), 12 (11), 1980-1992CODEN: RNARFU; ISSN:1355-8382. (Cold Spring Harbor Laboratory Press)The Ll.LtrB intron, from the low G+C gram-pos. bacterium Lactococcus lactis, was the first bacterial group II intron shown to splice and mobilize in vivo. The detailed retrohoming and retrotransposition pathways of Ll.LtrB were studied in both L. lactis and Escherichia coli. This bacterial retroelement has many features that would make it a good gene delivery vector. Here we report that the mobility efficiency of Ll.LtrB expressing LtrA in trans is only slightly affected by the insertion of fragments <100 nucleotides within the loop region of domain IV. In contrast, Ll.LtrB mobility efficiency is drastically decreased by the insertion of foreign sequences >1 kb. We demonstrate that the inhibitory effect caused by the addn. of expression cassettes on Ll.LtrB mobility efficiency is not sequence specific, and not due to the expression, or the toxicity, of the cargo genes. Using genetic screens, we demonstrate that in order to maintain intron mobility, the loop region of domain IV, more specifically domain IVb, is by far the best region to insert foreign sequences within Ll.LtrB. Poisoned primer extension and Northern blot analyses reveal that Ll.LtrB constructs harboring cargo sequences splice less efficiently, and show a significant redn. in lariat accumulation in L. lactis. This suggests that cargo-contg. Ll.LtrB variants are less stable. These results reveal the potential, yet limitations, of the Ll.LtrB group II intron to be used as a gene delivery vector, and validate the random insertion approach described in this study to create cargo-contg. Ll.LtrB variants that are mobile.
- 28Rawsthorne, H.; Turner, K. N.; Mills, D. A. Multicopy integration of heterologous genes, using the lactococcal group II intron targeted to bacterial insertion sequences. Appl. Environ. Microbiol. 2006, 72, 6088– 6093, DOI: 10.1128/AEM.02992-05Google Scholar28https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XpvVKjt70%253D&md5=c1d1e67be17b38c6d7a699b3b11ca654Multicopy integration of heterologous genes, using the lactococcal group II intron targeted to bacterial insertion sequencesRawsthorne, Helen; Turner, Kevin N.; Mills, David A.Applied and Environmental Microbiology (2006), 72 (9), 6088-6093CODEN: AEMIDF; ISSN:0099-2240. (American Society for Microbiology)Group II introns are mobile genetic elements that can be redirected to invade specific genes. Here the authors describe the use of the lactococcal group II intron, Ll.ltrB, to achieve multicopy delivery of heterologous genes into the genome of Lactococcus lactis IL1403-UCD without the need for selectable markers. Ll.ltrB was retargeted to invade three transposase genes, the tra gene found in IS904 (tra904), tra981, and tra983, of which 9, 10, and 14 copies, resp., were present in IL1403-UCD. Intron invasion of tra904, tra981, and tra983 allele groups occurred at high frequencies, and individual segregants possessed anywhere from one to nine copies of intron in the resp. tra alleles. To achieve multicopy delivery of a heterologous gene, a green fluorescent protein (GFP) marker was cloned into the tra904-targeted Ll.ltrB, and the resultant intron (Ll.ltrB::GFP) was induced to invade the L. lactis tra904 alleles. Segregants possessing Ll.ltrB::GFP in three, four, five, six, seven, and eight copies in different tra904 alleles were obtained. In general, increasing the chromosomal copy no. of Ll.ltrB::GFP resulted in strains expressing successively higher levels of GFP. However, strains possessing the same no. of Ll.ltrB::GFP copies within different sets of tra904 alleles exhibited differential GFP expression, and segregants possessing seven or eight copies of Ll.ltrB::GFP grew poorly upon induction, suggesting that GFP expression from certain combinations of alleles was detrimental. The highest level of GFP expression was obsd. from a specific six-copy variant that produced GFP at a level analogous to that obtained with a multicopy plasmid. In addn., the high level of GFP expression was stable for over 120 generations. This work demonstrates that stable multicopy integration of heterologous genes can be readily achieved in bacterial genomes with group II intron delivery by targeting repeated elements.
- 29Yao, J.; Zhong, J.; Fang, Y.; Geisinger, E.; Novick, R. P.; Lambowitz, A. M. Use of targetrons to disrupt essential and nonessential genes in Staphylococcus aureus reveals temperature sensitivity of Ll.LtrB group II intron splicing. RNA 2006, 12, 1271– 1281, DOI: 10.1261/rna.68706Google Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xms1Cqu78%253D&md5=4e089632c8b7390ccce87c3513cbe723Use of targetrons to disrupt essential and nonessential genes in Staphylococcus aureus reveals temperature sensitivity of Ll.LtrB group II intron splicingYao, Jun; Zhong, Jin; Fang, Yuan; Geisinger, Edward; Novick, Richard P.; Lambowitz, Alan M.RNA (2006), 12 (7), 1271-1281CODEN: RNARFU; ISSN:1355-8382. (Cold Spring Harbor Laboratory Press)We show that a targetron based on the Lactococcus lactis Ll.LtrB group II intron can be used for efficient chromosomal gene disruption in the human pathogen Staphylococcus aureus. Targetrons expressed from derivs. of vector pCN37, which uses a cadmium-inducible promoter, or pCN39, a deriv. of pCN37 with a temp.-sensitive replicon, gave site-specific disruptants of the hsa and seb genes in 37%-100% of plated colonies without selection. To disrupt hsa, an essential gene, we used a group II intron that integrates in the sense orientation relative to target gene transcription and thus could be removed by RNA splicing, enabling the prodn. of functional HSa protein. We show that because splicing of the Ll.LtrB intron by the intron-encoded protein is temp.-sensitive, this method yields a conditional hsa disruptant that grows at 32° but not 43°. The temp. sensitivity of the splicing reaction suggests a general means of obtaining one-step conditional disruptions in any organism. In nature, temp. sensitivity of group II intron splicing could limit the temp. range of an organism contg. a group II intron inserted in an essential gene.
- 30Karberg, M.; Guo, H.; Zhong, J.; Coon, R.; Perutka, J.; Lambowitz, A. M. Group II introns as controllable gene targeting vectors for genetic manipulation of bacteria. Nat. Biotechnol. 2001, 19, 1162– 1167, DOI: 10.1038/nbt1201-1162Google Scholar30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXptV2qsrw%253D&md5=3dad957336fc3ad3a37f4316d019bf4aGroup II introns as controllable gene targeting vectors for genetic manipulation of bacteriaKarberg, Michael; Guo, Huatao; Zhong, Jin; Coon, Robert; Perutka, Jiri; Lambowitz, Alan M.Nature Biotechnology (2001), 19 (12), 1162-1167CODEN: NABIF9; ISSN:1087-0156. (Nature America Inc.)Mobile group II introns can be retargeted to insert into virtually any desired DNA target. Here the authors show that retargeted group II introns can be used for highly specific chromosomal gene disruption in Escherichia coli and other bacteria at frequencies of 0.1-22%. Furthermore, the introns can be used to introduce targeted chromosomal breaks, which can be repaired by transformation with a homologous DNA fragment, enabling the introduction of point mutations. Because of their wide host range, mobile group II introns should be useful for genetic engineering and functional genomics in a wide variety of bacteria.
- 31García-Rodríguez, F. M.; Hernańdez-Gutiérrez, T.; Diáz-Prado, V.; Toro, N. Use of the computer-retargeted group II intron RmInt1 of Sinorhizobium meliloti for gene targeting. RNA Biol. 2014, 11, 391– 401, DOI: 10.4161/rna.28373Google Scholar31https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC2crlsFehsA%253D%253D&md5=49269dde7cb61dfd2adf91c2aa0d3971Use of the computer-retargeted group II intron RmInt1 of Sinorhizobium meliloti for gene targetingGarcia-Rodriguez Fernando M; Hernandez-Gutierrez Teresa; Diaz-Prado Vanessa; Toro NicolasRNA biology (2014), 11 (4), 391-401 ISSN:.Gene-targeting vectors derived from mobile group II introns capable of forming a ribonucleoprotein (RNP) complex containing excised intron lariat RNA and an intron-encoded protein (IEP) with reverse transcriptase (RT), maturase, and endonuclease (En) activities have been described. RmInt1 is an efficient mobile group II intron with an IEP lacking the En domain. We performed a comprehensive study of the rules governing RmInt1 target site recognition based on selection experiments with donor and recipient plasmid libraries, with randomization of the elements of the intron RNA involved in target recognition and the wild-type target site. The data obtained were used to develop a computer algorithm for identifying potential RmInt1 targets in any DNA sequence. Using this algorithm, we modified RmInt1 for the efficient recognition of DNA target sites at different locations in the Sinorhizobium meliloti chromosome. The retargeted RmInt1 integrated efficiently into the chromosome, regardless of the location of the target gene. Our results suggest that RmInt1 could be efficiently adapted for gene targeting.
- 32Cousineau, B.; Smith, D.; Lawrence-Cavanagh, S.; Mueller, J. E.; Yang, J.; Mills, D.; Manias, D.; Dunny, G.; Lambowitz, A. M.; Belfort, M. Retrohoming of a bacterial group II intron: Mobility via complete reverse splicing, independent of homologous DNA recombination. Cell 1998, 94, 451– 462, DOI: 10.1016/S0092-8674(00)81586-XGoogle Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXlslOktb8%253D&md5=ebbed8461cc1dadf9c870f01354bac6dRetrohoming of a bacterial group II intron: mobility via complete reverse splicing, independent of homologous DNA recombinationCousineau, Benoit; Smith, Dorie; Lawrence-Cavanagh, Stacey; Mueller, John E.; Yang, Jian; Mills, David; Dunny, Gary; Lambowitz, Alan M.; Belfort, MarleneCell (Cambridge, Massachusetts) (1998), 94 (4), 451-462CODEN: CELLB5; ISSN:0092-8674. (Cell Press)The mobile group II intron of Lactococcus lactis, LI.LtrB, provides the opportunity to analyze the homing pathway in genetically tractable bacterial systems. Here, we show that LI.LtrB mobility occurs by an RNA-based retrohoming mechanism in both Escherichia coli and L. lactis. Surprisingly, retrohoming occurs efficiently in the absence of RecA function, with a relaxed requirement for flanking exon homol. and without coconversion of exon markers. These results lead to a model for bacterial retrohoming in which the intron integrates into recipient DNA by complete reverse splicing and serves as the template for cDNA synthesis. The retrohoming reaction is completed in unprecedented fashion by a DNA repair event that is independent of homologous recombination between the alleles. Thus, LI.LtrB has many features of retrotransposons, with practical and evolutionary implications.
- 33Velázquez, E.; Lorenzo, V. D.; Al-Ramahi, Y. Recombination-Independent Genome Editing through CRISPR/Cas9-Enhanced TargeTron Delivery. ACS Synth. Biol. 2019, 8, 2186– 2193, DOI: 10.1021/acssynbio.9b00293Google Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhsF2ntbzN&md5=f06edd98e918993b500d50caa1978a9cRecombination-Independent Genome Editing through CRISPR/Cas9-Enhanced TargeTron DeliveryVelazquez, Elena; Lorenzo, Victor de; Al-Ramahi, YamalACS Synthetic Biology (2019), 8 (9), 2186-2193CODEN: ASBCD6; ISSN:2161-5063. (American Chemical Society)Group II introns were developed time ago as tools for the construction of knockout mutants in a wide range of organisms, ranging from Gram-pos. and Gram-neg. bacteria to human cells. Utilizing these introns is advantageous because they are independent of the host's DNA recombination machinery, they can carry heterologous sequences (and thus be used as vehicles for gene delivery), and they can be easily retargeted for subsequent insertions of addnl. genes at the user's will. Alas, the use of this platform has been limited, as insertion efficiencies greatly change depending on the target sites and cannot be predicted a priori. Moreover, the ability of introns to perform their own splicing and integration is compromised when they carry foreign sequences. To overcome these limitations, we merged the group II intron-based TargeTron system with CRISPR/Cas9 counter-selection. To this end, we first engineered a new group-II intron by replacing the retrotransposition-activated selectable marker (RAM) with ura3 and retargeting it to a new site in the lacZ gene of E. coli. Then, we showed proved that directing CRISPR/Cas9 towards the wild-type sequences dramatically increased the chances of finding clones that integrated the retrointron into the target lacZ sequence. The CRISPR-Cas9 counter selection strategy presented herein thus overcomes a major limitation that has prevented the use of group II introns as devices for gene delivery and genome editing at large in a recombination-independent fashion.
- 34Barrangou, R.; Fremaux, C.; Deveau, H.; Richards, M.; Boyaval, P.; Moineau, S.; Romero, D. A.; Horvath, P. CRISPR provides acquired resistance against viruses in prokaryotes. Science 2007, 315, 1709– 1712, DOI: 10.1126/science.1138140Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXjtlWntb8%253D&md5=b88c0100d2c0469213afda20e47c39cdCRISPR Provides Acquired Resistance Against Viruses in ProkaryotesBarrangou, Rodolphe; Fremaux, Christophe; Deveau, Helene; Richards, Melissa; Boyaval, Patrick; Moineau, Sylvain; Romero, Dennis A.; Horvath, PhilippeScience (Washington, DC, United States) (2007), 315 (5819), 1709-1712CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)Clustered regularly interspaced short palindromic repeats (CRISPR) are a distinctive feature of the genomes of most Bacteria and Archaea and are thought to be involved in resistance to bacteriophages. We found that, after viral challenge, bacteria integrated new spacers derived from phage genomic sequences. Removal or addn. of particular spacers modified the phage-resistance phenotype of the cell. Thus, CRISPR, together with assocd. cas genes, provided resistance against phages, and resistance specificity is detd. by spacer-phage sequence similarity.
- 35Jiang, W.; Bikard, D.; Cox, D.; Zhang, F.; Marraffini, L. A. RNA-guided editing of bacterial genomes using CRISPR-Cas systems. Nat. Biotechnol. 2013, 31, 233– 239, DOI: 10.1038/nbt.2508Google Scholar35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhsFCkurY%253D&md5=ea3bbfda133b51e71b114de2c9ceabfdRNA-guided editing of bacterial genomes using CRISPR-Cas systemsJiang, Wenyan; Bikard, David; Cox, David; Zhang, Feng; Marraffini, Luciano A.Nature Biotechnology (2013), 31 (3), 233-239CODEN: NABIF9; ISSN:1087-0156. (Nature Publishing Group)Here we use the clustered, regularly interspaced, short palindromic repeats (CRISPR)-assocd. Cas9 endonuclease complexed with dual-RNAs to introduce precise mutations in the genomes of Streptococcus pneumoniae and Escherichia coli. The approach relies on dual-RNA:Cas9-directed cleavage at the targeted genomic site to kill unmutated cells and circumvents the need for selectable markers or counter-selection systems. We reprogram dual-RNA:Cas9 specificity by changing the sequence of short CRISPR RNA (crRNA) to make single- and multinucleotide changes carried on editing templates. Simultaneous use of two crRNAs enables multiplex mutagenesis. In S. pneumoniae, nearly 100% of cells that were recovered using our approach contained the desired mutation, and in E. coli, 65% that were recovered contained the mutation, when the approach was used in combination with recombineering. We exhaustively analyze dual-RNA:Cas9 target requirements to define the range of targetable sequences and show strategies for editing sites that do not meet these requirements, suggesting the versatility of this technique for bacterial genome engineering.
- 36Aparicio, T.; de Lorenzo, V.; Martínez-García, E. CRISPR/Cas9-Based Counterselection Boosts Recombineering Efficiency in Pseudomonas putida. Biotechnol. J. 2018, 13, 1700161 DOI: 10.1002/biot.201700161Google ScholarThere is no corresponding record for this reference.
- 37Benedetti, I.; Nikel, P. I.; de Lorenzo, V. Data on the standardization of a cyclohexanone-responsive expression system for Gram-negative bacteria. Data Brief 2016, 6, 738– 744, DOI: 10.1016/j.dib.2016.01.022Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC28jgt1Onug%253D%253D&md5=5c2853ffbcc263263df26ad415e9484aData on the standardization of a cyclohexanone-responsive expression system for Gram-negative bacteriaBenedetti Ilaria; Nikel Pablo I; de Lorenzo VictorData in brief (2016), 6 (), 738-44 ISSN:2352-3409.Engineering of robust microbial cell factories requires the use of dedicated genetic tools somewhat different from those traditionally used for laboratory-adapted microorganisms. We have edited and formatted the ChnR/P chnB regulatory node of Acinetobacter johnsonii to ease the targeted engineering of ectopic gene expression in Gram-negative bacteria. The proposed compositional standard was thoroughly verified with a monomeric and superfolder green fluorescent protein (msf•GFP) in Escherichia coli. The expression data presented reflect a tightly controlled transcription initiation signal in response to cyclohexanone. Data in this article are related to the research paper "Genetic programming of catalytic Pseudomonas putida biofilms for boosting biodegradation of haloalkanes" [1].
- 38Yao, J.; Lambowitz, A. M. Gene targeting in Gram-negative bacteria by use of a mobile group II intron (“targetron”) expressed from a broad-host-range vector. Appl. Environ. Microbiol. 2007, 73, 2735– 2743, DOI: 10.1128/AEM.02829-06Google Scholar38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXkvFeqsbc%253D&md5=136a1d18e1873775be03c8940174d9feGene targeting in gram-negative bacteria by use of a mobile group II intron ("targetron") expressed from a broad-host-range vectorYao, Jun; Lambowitz, Alan M.Applied and Environmental Microbiology (2007), 73 (8), 2735-2743CODEN: AEMIDF; ISSN:0099-2240. (American Society for Microbiology)Mobile group II introns ("targetrons") can be programmed for insertion into virtually any desired DNA target with high frequency and specificity. Here, we show that targetrons expressed via an m-toluic acid-inducible promoter from a broad-host-range vector contg. an RK2 minireplicon can be used for efficient gene targeting in a variety of gram-neg. bacteria, including Escherichia coli, Pseudomonas aeruginosa, and Agrobacterium tumefaciens. Targetrons expressed from donor plasmids introduced by electroporation or conjugation yielded targeted disruptions at frequencies of 1 to 58% of screened colonies in the E. coli lacZ, P. aeruginosa pqsA and pqsH, and A. tumefaciens aopB and chvI genes. The development of this broad-host-range system for targetron expression should facilitate gene targeting in many bacteria.
- 39Martínez-García, E.; Aparicio, T.; Goñi-Moreno, A.; Fraile, S.; De Lorenzo, V. SEVA 2.0: An update of the Standard European Vector Architecture for de-/re-construction of bacterial functionalities. Nucleic Acids Res. 2015, 43, D1183– D1189, DOI: 10.1093/nar/gku1114Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhtVymtLfL&md5=8152ca1366af4c0fe1d99435b137bbcbSEVA 2.0: an update of the standard European vector architecture for de-/re-construction of bacterial functionalitiesMartinez-Garcia, Esteban; Aparicio, Tomas; Goni-Moreno, Angel; Fraile, Sofia; de Lorenzo, VictorNucleic Acids Research (2015), 43 (D1), D1183-D1189CODEN: NARHAD; ISSN:0305-1048. (Oxford University Press)A review. The Std. European Vector Architecture 2.0 database (SEVA-DB 2.0) is an improved and expanded version of the platform released in 2013 aimed at assisting the choice of optimal genetic tools for deconstructing and re-constructing complex prokaryotic phenotypes. By adopting simple compositional rules, the SEVA std. facilitates combinations of functional DNA segments that ease both the anal. and the engineering of diverse Gram-neg. bacteria for fundamental or biotechnol. purposes. The large no. of users of the SEVA-DB during its first two years of existence has resulted in a valuable feedback that we have exploited for fixing DNA sequence errors, improving the nomenclature of the SEVA plasmids, expanding the vector collection, adding new features to the web interface and encouraging contributions of materials from the community of users. The SEVA platform is also adopting the Synthetic Biol. Open Language (SBOL) for electronic-like description of the constructs available in the collection and their interfacing with genetic devices developed by other Synthetic Biol. communities. We advocate the SEVA format as one interim asset for the ongoing transition of genetic design of microorganisms from being a trial-and-error endeavor to become an authentic engineering discipline.
- 40Martínez-García, E.; Goñi-Moreno, A.; Bartley, B.; McLaughlin, J.; Sánchez-Sampedro, L.; Pascual Del Pozo, H.; Prieto Hernández, C.; Marletta, A. S.; De Lucrezia, D.; Sánchez-Fernández, G.; Fraile, S.; De Lorenzo, V. SEVA 3.0: An update of the Standard European Vector Architecture for enabling portability of genetic constructs among diverse bacterial hosts. Nucleic Acids Res. 2020, 48, D1164– D1170, DOI: 10.1093/nar/gkz1024Google Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhs1GltrbP&md5=4f5a11f9bf0a49d09a3ff86bcfdd1252SEVA 3.0: an update of the standard European vector architecture for enabling portability of genetic constructs among diverse bacterial hostsMartinez-Garcia, Esteban; Goni-Moreno, Angel; Bartley, Bryan; McLaughlin, James; Sanchez-Sampedro, Lucas; del Pozo, Hector Pascual; Hernandez, Clara Prieto; Marletta, Ada Serena; De Lucrezia, Davide; Sanchez-Fernandez, Guzman; Fraile, Sofia; de Lorenzo, VictorNucleic Acids Research (2020), 48 (D1), D1164-D1170CODEN: NARHAD; ISSN:1362-4962. (Oxford University Press)A review. The Std. European Vector Architecture 3.0 database is the update of the platform launched in 2013 both as a web-based resource and as a material repository of formatted genetic tools (mostly plasmids) for anal., construction and deployment of complex bacterial phenotypes. The period between the first version of SEVA-DB and the present time has witnessed several tech., computational and conceptual advances in genetic/genomic engineering of prokaryotes that have enabled upgrading of the utilities of the updated database. Novelties include not only a more user-friendly web interface and many more plasmid vectors, but also new links of the plasmids to advanced bioinformatic tools. These provide an intuitive visualization of the constructs at stake and a range of virtual manipulations of DNA segments that were not possible before. Finally, the list of canonical SEVA plasmids is available in machine-readable SBOL (Synthetic Biol. Open Language) format. This ensures interoperability with other platforms and affords simulations of their behavior under different in vivo conditions. We argue that the SEVA-DB will remain a useful resource for extending Synthetic Biol. approaches towards non-std. bacterial species as well as genetically programming new prokaryotic chassis for a suite of fundamental and biotechnol. endeavours.
- 41Zhong, J.; Karberg, M.; Lambowitz, A. M. Targeted and random bacterial gene disruption using a group II intron (targetron) vector containing a retrotransposition-activated selectable marker. Nucleic Acids Res. 2003, 31, 1656– 1664, DOI: 10.1093/nar/gkg248Google Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXit1yks7g%253D&md5=9ddea609f3357b85ce578dbb1967c657Targeted and random bacterial gene disruption using a group II intron (targetron) vector containing a retrotransposition-activated selectable markerZhong, Jin; Karberg, Michael; Lambowitz, Alan M.Nucleic Acids Research (2003), 31 (6), 1656-1664CODEN: NARHAD; ISSN:0305-1048. (Oxford University Press)Mobile group II introns have been used to develop a novel class of gene targeting vectors, targetrons, which employ base pairing for DNA target recognition and can thus be programmed to insert into any desired target DNA. Here, we have developed a targetron contg. a retrotransposition-activated selectable marker (RAM), which enables one-step bacterial gene disruption at near 100% efficiency after selection. The targetron can be generated via PCR without cloning, and after intron integration, the marker gene can be excised by recombination between flanking Flp recombinase sites, enabling multiple sequential disruptions. We also show that a RAM-targetron with randomized target site recognition sequences yields single insertions throughout the Escherichia coli genome, creating a gene knockout library. Anal. of the randomly selected insertion sites provides further insight into group II intron target site recognition rules. It also suggests that a subset of retrohoming events may occur by using a primer generated during DNA replication, and reveals a previously unsuspected bias for group II intron insertion near the chromosome replication origin. This insertional bias likely reflects at least in part the higher copy no. of origin proximal genes, but interaction with the replication machinery or other features of DNA structure or packaging may also contribute.
- 42Studier, F. W.; Moffatt, B. A. Use of bacteriophage T7 RNA polymerase to direct selective high-level expression of cloned genes. J. Mol. Biol. 1986, 189, 113– 130, DOI: 10.1016/0022-2836(86)90385-2Google Scholar42https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL28XktlKrsr4%253D&md5=3219300bc2f640fe9830c7518eb99bfcUse of bacteriophage T7 RNA polymerase to direct selective high-level expression of cloned genesStudier, F. William; Moffatt, Barbara A.Journal of Molecular Biology (1986), 189 (1), 113-30CODEN: JMOBAK; ISSN:0022-2836.A gene expression system based on phage T7 RNA polymerase [9014-24-8] was developed. T7 RNA polymerase is highly selective for its own promoters, which do not occur naturally in Escherichia coli. A relatively small amt. of T7 RNA polymerase provided from a cloned copy of T7 gene 1 is sufficient to direct high-level transcription from a T7 promoter in a multicopy plasmid. Such transcription can proceed several times around the plasmid without terminating and can be so active that transcription by E. coli RNA polymerase is greatly decreased. When a cleavage site for RNase III is introduced, discrete RNAs of plasmid length can accumulate. The natural transcription terminator from T7 DNA also works effectively in the plasmid. Both the rate of synthesis and the accumulation of RNA directed by T7 RNA polymerase can reach levels comparable with those for rRNAs in a normal cell. These high levels of accumulation suggest that the RNAs are relatively stable, perhaps in part because their great length and(or) stem-and-loop structures at their 3' ends help to protect them against exonucleolytic degrdn. Apparently, a specific mRNA produced by T7 RNA polymerase can rapidly sat. the translational machinery of E. coli, so that the rate of protein synthesis from such an mRNA will depend primarily on the efficiency of its translation. When the mRNA is efficiently translated, a target protein can accumulate to >50% of the total cell protein in ≤3 h. Two ways were used to deliver active T7 RNA polymerase to the cell: (1) infection by a λ deriv. that carried gene 1; or (2) induction of a chromosomal copy of gene 1 under control of the lacUV5 promoter. When gene 1 is delivered by infection, very toxic target genes can be maintained silently in the cell until T7 RNA polymerase is introduced, when they rapidly become expressed at high levels. When gene 1 is resident in the chromosome, even the very low basal levels of T7 RNA polymerase present in the uninduced cell can prevent the establishment of plasmids carrying toxic target genes, or make the plasmid unstable. But if the target plasmid can be maintained, induction of chromosomal gene 1 can be a convenient way to produce large amts. of target RNA and(or) protein. T7 RNA polymerase seems to be capable of transcribing almost any DNA linked to a T7 promoter, so the T7 expression system should be capable of transcribing almost any gene or its complement in E. coli. Comparable T7 expression systems can be developed in other types of cells.
- 43Dubendorf, J. W.; Studier, F. W. Controlling basal expression in an inducible T7 expression system by blocking the target T7 promoter with lac repressor. J. Mol. Biol. 1991, 219, 45– 59, DOI: 10.1016/0022-2836(91)90856-2Google ScholarThere is no corresponding record for this reference.
- 44Angius, F.; Ilioaia, O.; Amrani, A.; Suisse, A.; Rosset, L.; Legrand, A.; Abou-Hamdan, A.; Uzan, M.; Zito, F.; Miroux, B. A novel regulation mechanism of the T7 RNA polymerase based expression system improves overproduction and folding of membrane proteins. Sci. Rep. 2018, 8, 8572 DOI: 10.1038/s41598-018-26668-yGoogle Scholar44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC1Mbkt12isg%253D%253D&md5=0256960f3095dbbacbc5be9d233b3c38A novel regulation mechanism of the T7 RNA polymerase based expression system improves overproduction and folding of membrane proteinsAngius Federica; Ilioaia Oana; Amrani Amira; Suisse Annabelle; Rosset Lindsay; Legrand Amelie; Abou-Hamdan Abbas; Uzan Marc; Zito Francesca; Miroux Bruno; Suisse Annabelle; Abou-Hamdan AbbasScientific reports (2018), 8 (1), 8572 ISSN:.Membrane protein (MP) overproduction is one of the major bottlenecks in structural genomics and biotechnology. Despite the emergence of eukaryotic expression systems, bacteria remain a cost effective and powerful tool for protein production. The T7 RNA polymerase (T7RNAP)-based expression system is a successful and efficient expression system, which achieves high-level production of proteins. However some foreign MPs require a fine-tuning of their expression to minimize the toxicity associated with their production. Here we report a novel regulation mechanism for the T7 expression system. We have isolated two bacterial hosts, namely C44(DE3) and C45(DE3), harboring a stop codon in the T7RNAP gene, whose translation is under the control of the basal nonsense suppressive activity of the BL21(DE3) host. Evaluation of hosts with superfolder green fluorescent protein (sfGFP) revealed an unprecedented tighter control of transgene expression with a marked accumulation of the recombinant protein during stationary phase. Analysis of a collection of twenty MP fused to GFP showed an improved production yield and quality of several bacterial MPs and of one human monotopic MP. These mutant hosts are complementary to the other existing T7 hosts and will increase the versatility of the T7 expression system.
- 45Ichiyanagi, K.; Beauregard, A.; Lawrence, S.; Smith, D.; Cousineau, B.; Belfort, M. Retrotransposition of the Ll.LtrB group II intron proceeds predominantly via reverse splicing into DNA targets. Mol. Microbiol. 2002, 46, 1259– 1272, DOI: 10.1046/j.1365-2958.2002.03226.xGoogle Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XpsFGisL0%253D&md5=a588926c206e0bf19be2da6bb66a36c9Retrotransposition of the LI.LtrB group II intron proceeds predominantly via reverse splicing into DNA targetsIchiyanagi, Kenji; Beauregard, Arthur; Lawrence, Stacey; Smith, Dorie; Cousineau, Benoit; Belfort, MarleneMolecular Microbiology (2002), 46 (5), 1259-1272CODEN: MOMIEE; ISSN:0950-382X. (Blackwell Science Ltd.)Catalytic group II introns are mobile retroelements that invade cognate intronless genes via retrohoming, where the introns reverse splice into double-stranded DNA (dsDNA) targets. They can also retrotranspose to ectopic sites at low frequencies. Whereas our previous studies with a bacterial intron, LI.LtrB, supported frequent use of RNA targets during retrotransposition, recent expts. with a retrotransposition indicator gene indicate that DNA, rather than RNA, is a prominent target, with both dsDNA and single-stranded DNA (ssDNA) as possibilities. Thus retrotransposition occurs in both transcriptional sense and antisense orientations of target genes, and is largely independent of homologous DNA recombination and of the endonuclease function of the intron-encoded protein, LtrA. Models based on both dsDNA and ssDNA targeting are presented. Interestingly, retrotransposition is biased toward the template for lagging-strand DNA synthesis, which suggests the possibility of the replication folk as a source of ssDNA. Consistent with some use of ssDNA targets, many retrotransposition sites lack nucleotides crit. for the unwinding of target duplex DNA. Moreover, in vitro the intron reverse spliced into ssDNA more efficiently than dsDNA substrates for some of the retrotransposition sites. Furthermore, many bacterial group II introns reside on the lagging-strand template, hinting at a role for DNA replication in intron dispersal in nature.
- 46Dickson, L.; Huang, H. R.; Liu, L.; Matsuura, M.; Lambowitz, A. M.; Perlman, P. S. Retrotransposition of a yeast group II intron occurs by reverse splicing directly into ectopic DNA sites. Proc. Natl. Acad. Sci. U.S.A. 2001, 98, 13207– 13212, DOI: 10.1073/pnas.231494498Google Scholar46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXosFygtrs%253D&md5=5bd9e382b0d94d8b6bb91f15232e59ccRetrotransposition of a yeast group II intron occurs by reverse splicing directly into ectopic DNA sitesDickson, Lorna; Huang, Hon-Ren; Liu, Lu; Matsuura, Manabu; Lambowitz, Alan M.; Perlman, Philip S.Proceedings of the National Academy of Sciences of the United States of America (2001), 98 (23), 13207-13212CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)Group II introns, the presumed ancestors of nuclear pre-mRNA introns, are site-specific retroelements. In addn. to "homing" to unoccupied sites in intronless alleles, group II introns transpose at low frequency to ectopic sites that resemble the normal homing site. Two general mechanisms have been proposed for group II intron transposition, one involving reverse splicing of the intron RNA directly into an ectopic DNA site, and the other involving reverse splicing into a site in RNA followed by reverse transcription and integration of the resulting cDNA by homologous recombination. Here, by using an "inverted-site" strategy, we show that the yeast mtDNA group II intron aI1 retrotransposes by reverse splicing directly into an ectopic DNA site. This same mechanism could account for other previously described ectopic transposition events in fungi and bacteria and may have contributed to the dispersal of group II introns into different genes.
- 47Coros, C. J.; Landthaler, M.; Piazza, C. L.; Beauregard, A.; Esposito, D.; Perutka, J.; Lambowitz, A. M.; Belfort, M. Retrotransposition strategies of the Lactococcus lactis Ll.LtrB group II intron are dictated by host identity and cellular environment. Mol. Microbiol. 2005, 56, 509– 524, DOI: 10.1111/j.1365-2958.2005.04554.xGoogle Scholar47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXjsFWmtLs%253D&md5=d68f89a37bc9cd4bd4c9c8f403dd2fd5Retrotransposition strategies of the lactococcus lactis Ll.LtrB group II intron are dictated by host identity and cellular environmentCoros, Colin J.; Landthaler, Markus; Piazza, Carol Lyn; Beauregard, Arthur; Esposito, Donna; Perutka, Jiri; Lambowitz, Alan M.; Belfort, MarleneMolecular Microbiology (2005), 56 (2), 509-524CODEN: MOMIEE; ISSN:0950-382X. (Blackwell Publishing Ltd.)Group II introns are mobile retroelements that invade their cognate intron-minus gene in a process known as retrohoming. They can also retrotranspose to ectopic sites at low frequency. Previous studies of the Lactococcus lactis intron Ll.LtrB indicated that in its native host, as in Escherichia coli, retrohoming occurs by the intron RNA reverse splicing into double-stranded DNA (dsDNA) through an endonuclease-dependent pathway. However, in retrotransposition in L. lactis, the intron inserts predominantly into single-stranded DNA (ssDNA), in an endonuclease-independent manner. This work describes the retrotransposition of the Ll.LtrB intron in E. coli, using a retrotransposition indicator gene previously employed in our L. lactis studies. Unlike in L. lactis, in E. coli, Ll.LtrB retrotransposed frequently into dsDNA, and the process was dependent on the endonuclease activity of the intron-encoded protein. Further, the endonuclease-dependent insertions preferentially occurred around the origin and terminus of chromosomal DNA replication. Insertions in E. coli can also occur through an endonuclease-independent pathway, and, as in L. lactis, such events have a more random integration pattern. Together these findings show that Ll.LtrB can retrotranspose through at least two distinct mechanisms and that the host environment influences the choice of integration pathway. Addnl., growth conditions affect the insertion pattern. We propose a model in which DNA replication, compactness of the nucleoid and chromosomal localization influence target site preference.
- 48Flynn, P. J.; Reece, R. J. Activation of Transcription by Metabolic Intermediates of the Pyrimidine Biosynthetic Pathway. Mol. Cell. Biol. 1999, 19, 882– 888, DOI: 10.1128/MCB.19.1.882Google Scholar48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXhvFyitA%253D%253D&md5=c3a936b1eda8f038270317fa526b055eActivation of transcription by metabolic intermediates of the pyrimidine biosynthetic pathwayFlynn, Paul J.; Reece, Richard J.Molecular and Cellular Biology (1999), 19 (1), 882-888CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)Saccharomyces cerevisiae responds to pyrimidine starvation by increasing the expression of four URA genes, encoding the enzymes of de novo pyrimidine biosynthesis, three- to eightfold. The increase in gene expression is dependent on a transcriptional activator protein, Ppr1p. Here, we investigate the mechanism by which the transcriptional activity of Ppr1p responds to the level of pyrimidine biosynthetic intermediates. We find that purified Ppr1p is unable to promote activation of transcription in an in vitro system. Transcriptional activation by Ppr1p can be obsd., however, if either dihydroorotic acid (DHO) or orotic acid (OA) is included in the transcription reactions. The transcriptional activation function and the DHO/OA-responsive element of Ppr1p localize to the carboxyl-terminal 134 amino acids of the protein. Thus, Ppr1p directly senses the level of early pyrimidine biosynthetic intermediates within the cell and activates the expression of genes encoding proteins required later in the pathway. These results are discussed in terms of (i) regulation of the pyrimidine biosynthetic pathway and (ii) a novel mechanism of regulating gene expression.
- 49Boeke, J. D.; La Croute, F.; Fink, G. R. A positive selection for mutants lacking orotidine-5′-phosphate decarboxylase activity in yeast: 5-fluoro-orotic acid resistance. Mol. Gen. Genet. 1984, 197, 345– 346, DOI: 10.1007/BF00330984Google Scholar49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2MXhs1elsrg%253D&md5=1fbe754fa41c9f426317388cbc7381abA positive selection for mutants lacking orotidine-5'-phosphate decarboxylase activity in yeast: 5-fluoroorotic acid resistanceBoeke, Jef D.; LaCroute, Francois; Fink, Gerald R.Molecular and General Genetics (1984), 197 (2), 345-6CODEN: MGGEAE; ISSN:0026-8925.Mutations at the URA3 locus of Saccharomyces cerevisiae can be obtained by a pos. selection. Wild-type strains of yeast (or ura3 mutant strains contg. a plasmid-borne URA3+ gene) are unable to grow on medium contg. the pyrimidine analog 5-fluoroorotic acid, whereas ura3- mutants grow normally. This selection, based on the loss of orotidine-5'-phosphate decarboxylase activity seems applicable to a variety of eucaryotic and procaryotic cells.
- 50Galvão, T. C.; De Lorenzo, V. Adaptation of the yeast URA3 selection system to Gram-negative bacteria and generation of a ΔbetCDE Pseudomonas putida strain. Appl. Environ. Microbiol. 2005, 71, 883– 892, DOI: 10.1128/AEM.71.2.883-892.2005Google Scholar50https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXhsVKgtrw%253D&md5=03f764b6e6c5e72eb582e0cfcf77882bAdaptation of the yeast URA3 selection system to Gram-negative bacteria and generation of a ΔbetCDE Pseudomonas putida strainGalvao, Teca Calcagno; De Lorenzo, VictorApplied and Environmental Microbiology (2005), 71 (2), 883-892CODEN: AEMIDF; ISSN:0099-2240. (American Society for Microbiology)A general procedure for efficient generation of gene knockouts in gram-neg. bacteria by the adaptation of the Saccharomyces cerevisiae URA3 selection system is described. A Pseudomonas putida strain lacking the URA3 homolog pyrF (encoding orotidine-5'-phosphate decarboxylase) was constructed, allowing the use of a plasmid-borne copy of the gene as the target of selection. The delivery vector pTEC contains the pyrF gene and promoter, a conditional origin of replication (oriR6K), an origin of transfer (mobRK2), and an antibiotic selection marker flanked by multiple sites for cloning appropriate DNA segments. The versatility of pyrF as a selection system, allowing both pos. and neg. selection of the marker, and the robustness of the selection, where pyrF is assocd. with uracil prototrophy and fluoroorotic acid sensitivity, make this setup a powerful tool for efficient homologous gene replacement in gram-neg. bacteria. The system has been instrumental for complete deletion of the P. putida choline-O-sulfate utilization operon betCDE, a mutant which could not be produced by any of the other genetic strategies available.
- 51Cousineau, B.; Lawrence, S.; Smith, D.; Belfort, M. Retrotransposition of a bacterial group II intron. Nature 2000, 404, 1018– 1021, DOI: 10.1038/35010029Google Scholar51https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXjtFynu7s%253D&md5=8ce276b86062da0e7ec9423eae506dd0Retrotransposition of a bacterial group II intronCousineau, Benoit; Lawrence, Stacey; Smith, Dorie; Belfort, MarleneNature (London) (2000), 404 (6781), 1018-1021CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)Self-splicing group II introns may be the evolutionary progenitors of eukaryotic spliceosomal introns, but the route by which they invade new chromosomal sites is unknown. To address the mechanism by which group II introns are disseminated, we have studied the bacterial L1.LtrB intron from Lactococcus lactis. The protein product of this intron, LtrA, possesses maturase, reverse transcriptase and endonuclease enzymic activities. Together with the intron, LtrA forms a ribonucleoprotein (RNP) complex which mediates a process known as retrohoming. In retrohoming, the intron reverse splices into a cognate intronless DNA site. Integration of a DNA copy of the intron is recombinase independent but requires all three activities of LtrA. Here we report the first exptl. demonstration of a group II intron invading ectopic chromosomal sites, which occurs by a distinct retrotransposition mechanism. This retrotransposition process is endonuclease-independent and recombinase-dependent, and is likely to involve reverse splicing of the intron RNA into cellular RNA targets. These retrotranspositions suggest a mechanism by which splicesomal introns may have become widely dispersed.
- 52Golomb, M.; Chamberlin, M. Characterization of T7 specific ribonucleic acid polymerase. IV. Resolution of the major in vitro transcripts by gel electrophoresis. J. Biol. Chem. 1974, 249, 2858– 2863, DOI: 10.1016/S0021-9258(19)42709-9Google Scholar52https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE2cXksVKrs70%253D&md5=5abe8d2ccb186b1d414061a838117d69Characterization of T7-specific ribonucleic acid polymerase. IV. Resolution of the major in vitro transcripts by gel electrophoresisGolomb, Miriam; Chamberlin, MichaelJournal of Biological Chemistry (1974), 249 (9), 2858-63CODEN: JBCHA3; ISSN:0021-9258.The major in vitro RNA transcripts synthesized by T7 RNA polymerase with T7 and T3 DNA templates were resolved by electrophoresis on polyacrylamide gels. Six discrete size classes of T7 RNAs were found, and designated I to VI. The apparent mol. wts., estd. from their electrophoretic mobilities, were 2 × 105-5 × 106. At least 2 minor RNA species with mol. wts. >5 × 106 were also detected. The 6 major T7 RNA species were synthesized in approx. equimolar amts., with the exception of species III, which was made in ∼ twice this amt. Perhaps, species III is a mixt. of two RNAs transcribed from sep. regions of the T7 genome. Hence, the 6 major T7 RNA species are tentatively identified with 7 late transcription units on the phage chromosome, which are read with equal efficiencies. The 6 (or 7) transcription products were inititated independently with GTP at the 5' terminus, and were elongated at a rate of 230 nucleotides/s under standard in vitro conditions. When T7 RNA polymerase was used to transcribe T3 DNA, a single RNA transcript was found, with a mol. wt. similar to that of T7 RNA species III. This suggests an explanation for the reduced rate of T3 RNA synthesis by T7 RNA polymerase in vitro, and implies that T3 promoter sites read by the T3 RNA polymerase are heterogeneous in nature.
- 53Chamberlin, M.; Ryan, T. 4 Bacteriophage DNA-Dependent RNA Polymerases. Enzymes 1982, 15, 87– 108Google Scholar53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL38XltVGqsLo%253D&md5=edd42f3f5bbcbd40064d5fdb469be1e4Bacteriophage DNA-dependent RNA polymerasesChamberlin, Michael J.; Ryan, T.(1982), 15 (Pt. B), 87-108CODEN: 25GLAS ISSN:. (Academic)A review with 87 refs.
- 54Benedetti, I.; de Lorenzo, V.; Nikel, P. I. Genetic programming of catalytic Pseudomonas putida biofilms for boosting biodegradation of haloalkanes. Metab. Eng. 2016, 33, 109– 118, DOI: 10.1016/j.ymben.2015.11.004Google Scholar54https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhvFems7rL&md5=64291830c898a5f4ecfb5fef5313598cGenetic programming of catalytic Pseudomonas putida biofilms for boosting biodegradation of haloalkanesBenedetti, Ilaria; de Lorenzo, Victor; Nikel, Pablo I.Metabolic Engineering (2016), 33 (), 109-118CODEN: MEENFM; ISSN:1096-7176. (Elsevier B. V.)Bacterial biofilms outperform planktonic counterparts in whole-cell biocatalysis. The transition between planktonic and biofilm lifestyles of the platform strain Pseudomonas putida KT2440 is ruled by a regulatory network controlling the levels of the trigger signal cyclic di-GMP (c-di-GMP). This circumstance was exploited for designing a genetic device that over-runs the synthesis or degrdn. of c-di-GMP, thus making P. putida to form biofilms at user's will. For this purpose, the transcription of either yedQ (diguanylate cyclase) or yhjH (c-di-GMP phosphodiesterase) from Escherichia coli was artificially placed under the tight control of a cyclohexanone-responsive expression system. The resulting strain was subsequently endowed with a synthetic operon and tested for 1-chlorobutane biodegrdn. Upon addn. of cyclohexanone to the culture medium, the thereby designed P. putida cells formed biofilms displaying high dehalogenase activity. These results show that the morphologies and phys. forms of whole-cell biocatalysts can be genetically programmed while purposely designing their biochem. activity.
- 55Martínez-Abarca, F.; García-Rodríguez, F. M.; Toro, N. Homing of a bacterial group II intron with an intron-encoded protein lacking a recognizable endonuclease domain. Mol. Microbiol. 2000, 35, 1405– 1412, DOI: 10.1046/j.1365-2958.2000.01804.xGoogle Scholar55https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXis1Wrtro%253D&md5=7cf3687672ca684154f02972891ab74bHoming of a bacterial group II intron with an intron-encoded protein lacking a recognizable endonuclease domainMartinez-Abarca, Francisco; Garcia-Rodriguez, Fernando M.; Toro, NicolasMolecular Microbiology (2000), 35 (6), 1405-1412CODEN: MOMIEE; ISSN:0950-382X. (Blackwell Science Ltd.)Rmlnt1 is a functional group II intron found in Sinorhizobium meliloti where it interrupts a group of IS elements of the IS630-Tc1 family. In contrast to many other group II introns, the intron-encoded protein (IEP) of Rmlnt1 lacks the characteristic conserved part of the Zn domain assocd. with the IEP endonuclease activity. Nevertheless, in this study, we show that Rmlnt1 is capable of inserting into a vector contg. the DNA spanning the Rmlnt1 target site from the genome of S. meliloti. Efficient homing was also obsd. in the absence of homologous recombination (RecA- strains). In addn., it is shown that Rmlnt1 is able to move to its target in a heterologous host (S. medicae). Homing of Rmlnt1 occurs very efficiently upon DNA target uptake (conjugation/electroporation) by the host cell resulting in a proportion of invaded target of 11-30%. Afterwards, the remaining intronless target DNA is protected from intron invasion.
- 56Cook, T. B.; Rand, J. M.; Nurani, W.; Courtney, D. K.; Liu, S. A.; Pfleger, B. F. Genetic tools for reliable gene expression and recombineering in Pseudomonas putida. J. Ind. Microbiol. Biotechnol. 2018, 45, 517– 527, DOI: 10.1007/s10295-017-2001-5Google Scholar56https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXjt1GltA%253D%253D&md5=8404ada292d1376d0fa2b08c54d1863aGenetic tools for reliable gene expression and recombineering in Pseudomonas putidaCook, Taylor B.; Rand, Jacqueline M.; Nurani, Wasti; Courtney, Dylan K.; Liu, Sophia A.; Pfleger, Brian F.Journal of Industrial Microbiology & Biotechnology (2018), 45 (7), 517-527CODEN: JIMBFL; ISSN:1367-5435. (Springer)Pseudomonas putida is a promising bacterial host for producing natural products, such as polyketides and nonribosomal peptides. In these types of projects, researchers need a genetic toolbox consisting of plasmids, characterized promoters, and techniques for rapidly editing the genome. Past reports described constitutive promoter libraries, a suite of broad host range plasmids that replicate in P. putida, and genome-editing methods. To augment those tools, we have characterized a set of inducible promoters and discovered that IPTG-inducible promoter systems have poor dynamic range due to overexpression of the LacI repressor. By replacing the promoter driving lacI expression with weaker promoters, we increased the fold induction of an IPTG-inducible promoter in P. putida KT2440 to 80-fold. Upon discovering that gene expression from a plasmid was unpredictable when using a high-copy mutant of the BBR1 origin, we detd. the copy nos. of several broad host range origins and found that plasmid copy nos. are significantly higher in P. putida KT2440 than in the synthetic biol. workhorse, Escherichia coli. Lastly, we developed a λRed/Cas9 recombineering method in P. putida KT2440 using the genetic tools that we characterized. This method enabled the creation of scarless mutations without the need for performing classic two-step integration and marker removal protocols that depend on selection and counterselection genes. With the method, we generated four scarless deletions, three of which we were unable to create using a previously established genome-editing technique.
- 57Bagdasarian, M.; Lurz, R.; Rückert, B.; Franklin, F. C. H.; Bagdasarian, M. M.; Frey, J.; Timmis, K. N. Specific-purpose plasmid cloning vectors II. Broad host range, high copy number, RSF 1010-derived vectors, and a host-vector system for gene cloning in Pseudomonas. Gene 1981, 16, 237– 247, DOI: 10.1016/0378-1119(81)90080-9Google Scholar57https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL38XhsFKlsrg%253D&md5=886ab175b0c338a0ec3030fe6e38c178Specific-purpose plasmid cloning vectors. II. Broad host range, high copy number, RSF1010-derived vectors, and a host-vector system for gene cloning in PseudomonasBagdasarian, M.; Lurz, R.; Rueckert, B.; Franklin, F. C. H.; Bagdasarian, M. M.; Frey, J.; Timmis, K. N.Gene (1981), 16 (1-3), 237-47CODEN: GENED6; ISSN:0378-1119.Host-vector systems were developed for gene cloning in the metabolically versatile bacterial genus Pseudomonas. They comprise restriction-neg. host strains of P. aeruginosa and P. putida and new cloning vectors derived from the high-copy-no., broad-host-range plasmid RSF1010, which are stably maintained in a wide range of gram-neg. bacteria. These plasmids contain EcoRI, SstI, HindIII, XmaI, XhoI, SalI, BamHI, and ClaI insertion sites. All cloning sites, except for BamHI and ClaI, are located within antibiotic-resistance genes; insertional inactivation of these genes during hybrid plasmid formation provides a readily scored phenotypic change for the rapid identification of bacterial clones carrying such hybrids. One of the new vector plasmids is a cosmid that may be used for the selective cloning of large DNA fragments by in vitro λ packaging. An analogous series of vectors that are defective in their plasmid-mobilization function, and that exhibit a degree of biol. containment comparable to that of current Escherichia coli vector plasmids, are also described.
- 58Jahn, M.; Vorpahl, C.; Hübschmann, T.; Harms, H.; Müller, S. Copy number variability of expression plasmids determined by cell sorting and Droplet Digital PCR. Microb. Cell Fact. 2016, 15, 211 DOI: 10.1186/s12934-016-0610-8Google Scholar58https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXjsFCrsbc%253D&md5=3245e40449919390aec4633a06832341Copy number variability of expression plasmids determined by cell sorting and Droplet Digital PCRJahn, Michael; Vorpahl, Carsten; Huebschmann, Thomas; Harms, Hauke; Mueller, SusannMicrobial Cell Factories (2016), 15 (), 211/1-211/12CODEN: MCFICT; ISSN:1475-2859. (BioMed Central Ltd.)Plasmids are widely used for mol. cloning or prodn. of proteins in lab. and industrial settings. Const. modification has brought forth countless plasmid vectors whose characteristics in terms of av. plasmid copy no. (PCN) and stability are rarely known. The crucial factor detg. the PCN is the replication system; most replication systems in use today belong to a small no. of different classes and are available through repositories like the Std. European Vector Architecture (SEVA). In this study, the PCN was detd. in a set of seven SEVA-based expression plasmids only differing in the replication system. The av. PCN for all constructs was detd. by Droplet Digital PCR and ranged between 2 and 40 per chromosome in the host organism Escherichia coli. Furthermore, a plasmid-encoded EGFP reporter protein served as a means to assess variability in reporter gene expression on the single cell level. Only cells with one type of plasmid (RSF1010 replication system) showed a high degree of heterogeneity with a clear bimodal distribution of EGFP intensity while the others showed a normal distribution. The heterogeneous RSF1010-carrying cell population and one normally distributed population (ColE1 replication system) were further analyzed by sorting cells of sub-populations selected according to EGFP intensity. For both plasmids, low and highly fluorescent sub-populations showed a remarkable difference in PCN, ranging from 9.2 to 123.4 for ColE1 and from 0.5 to 11.8 for RSF1010, resp. The av. PCN detd. here for a set of standardized plasmids was generally at the lower end of previously reported ranges and not related to the degree of heterogeneity. Further characterization of a heterogeneous and a homogeneous population demonstrated considerable differences in the PCN of sub-populations. We therefore present direct mol. evidence that the av. PCN does not represent the true no. of plasmid mols. in individual cells.
- 59Pyne, M. E.; Moo-Young, M.; Chung, D. A.; Chou, C. P. Coupling the CRISPR/Cas9 system with lambda red recombineering enables simplified chromosomal gene replacement in Escherichia coli. Appl. Environ. Microbiol. 2015, 81, 5103– 5114, DOI: 10.1128/AEM.01248-15Google Scholar59https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXht1Ojt7bP&md5=993a4d147fcb4c66368e6ec73bae3f7aCoupling the CRISPR/Cas9 system with lambda Red recombineering enables simplified chromosomal gene replacement in Escherichia coliPyne, Michael E.; Moo-Young, Murray; Chung, Duane A.; Chou, C. PerryApplied and Environmental Microbiology (2015), 81 (15), 5103-5114CODEN: AEMIDF; ISSN:1098-5336. (American Society for Microbiology)To date, most genetic engineering approaches coupling the type II Streptococcus pyogenes clustered regularly interspaced short palindromic repeat (CRISPR)/Cas9 system to lambda Red recombineering have involved minor single nucleotide mutations. Here we show that procedures for carrying out more complex chromosomal gene replacements in Escherichia coli can be substantially enhanced through implementation of CRISPR/Cas9 genome editing. We developed a three-plasmid approach that allows not only highly efficient recombination of short single-stranded oligonucleotides but also replacement of multigene chromosomal stretches of DNA with large PCR products. By systematically challenging the proposed system with respect to the magnitude of chromosomal deletion and size of DNA insertion, we demonstrated DNA deletions of up to 19.4 kb, encompassing 19 nonessential chromosomal genes, and insertion of up to 3 kb of heterologous DNA with recombination efficiencies permitting mutant detection by colony PCR screening. Since CRISPR/Cas9-coupled recombineering does not rely on the use of chromosome- encoded antibiotic resistance, or flippase recombination for antibiotic marker recycling, our approach is simpler, less labor-intensive, and allows efficient prodn. of gene replacement mutants that are both markerless and "scar"-less.
- 60Moreb, E. A.; Hoover, B.; Yaseen, A.; Valyasevi, N.; Roecker, Z.; Menacho-Melgar, R.; Lynch, M. D. Managing the SOS Response for Enhanced CRISPR-Cas-Based Recombineering in E. coli through Transient Inhibition of Host RecA Activity. ACS Synth. Biol. 2017, 6, 2209– 2218, DOI: 10.1021/acssynbio.7b00174Google Scholar60https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhsV2murrN&md5=70fcee7818b69824c249a54d3385fe39Managing the SOS Response for Enhanced CRISPR-Cas-Based Recombineering in E. coli through Transient Inhibition of Host RecA ActivityMoreb, Eirik Adim; Hoover, Benjamin; Yaseen, Adam; Valyasevi, Nisakorn; Roecker, Zoe; Menacho-Melgar, Romel; Lynch, Michael D.ACS Synthetic Biology (2017), 6 (12), 2209-2218CODEN: ASBCD6; ISSN:2161-5063. (American Chemical Society)Phage-derived "recombineering" methods are utilized for bacterial genome editing. Recombineering results in a heterogeneous population of modified and unmodified chromosomes, and therefore selection methods, such as CRISPR-Cas9, are required to select for edited clones. Cells can evade CRISPR-Cas-induced cell death through recA-mediated induction of the SOS response. The SOS response increases RecA dependent repair as well as mutation rates through induction of the umuDC error prone polymerase. As a result, CRISPR-Cas selection is more efficient in recA mutants. We report an approach to inhibiting the SOS response and RecA activity through the expression of a mutant dominant neg. form of RecA, which incorporates into wild type RecA filaments and inhibits activity. Using a plasmid-based system in which Cas9 and recA mutants are coexpressed, we can achieve increased efficiency and consistency of CRISPR-Cas9-mediated selection and recombineering in E. coli, while reducing the induction of the SOS response. To date, this approach has been shown to be independent of recA genotype and host strain lineage. Using this system, we demonstrate increased CRISPR-Cas selection efficacy with over 10,000 guides covering the E. coli chromosome. The use of dominant neg. RecA or homologues may be of broad use in bacterial CRISPR-Cas-based genome editing where the SOS pathways are present.
- 61Tellechea-Luzardo, J.; Winterhalter, C.; Widera, P.; Kozyra, J.; De Lorenzo, V.; Krasnogor, N. Linking Engineered Cells to Their Digital Twins: A Version Control System for Strain Engineering. ACS Synth. Biol. 2020, 9, 536– 545, DOI: 10.1021/acssynbio.9b00400Google Scholar61https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXjsVOntr0%253D&md5=282b32256dcc6a167abd1fb5b8484f29Linking Engineered Cells to Their Digital Twins: A Version Control System for Strain EngineeringTellechea-Luzardo, Jonathan; Winterhalter, Charles; Widera, Pawel; Kozyra, Jerzy; de Lorenzo, Victor; Krasnogor, NatalioACS Synthetic Biology (2020), 9 (3), 536-545CODEN: ASBCD6; ISSN:2161-5063. (American Chemical Society)As DNA sequencing and synthesis become cheaper and more easily accessible, the scale and complexity of biol. engineering projects is set to grow. Yet, although there is an accelerating convergence between biotechnol. and digital technol., a deficit in software and lab. techniques diminishes the ability to make biotechnol. more agile, reproducible and transparent while, at the same time, limiting the security and safety of synthetic biol. constructs. To partially address some of these problems, this paper presents an approach for phys. linking engineered cells to their digital footprint-we called it digital twinning. This enables the tracking of the entire engineering history of a cell line in a specialized version control system for collaborative strain engineering via simple barcoding protocols.
- 62Tellechea-Luzardo, J.; Hobbs, L.; Velázquez, E.; Pelechova, L.; Woods, S.; de Lorenzo, V.; Krasnogor, N. Versioning Biological Cells for Trustworthy Cell Engineering. bioRxiv 2021. DOI: 10.1101/2021.04.23.441106 .Google ScholarThere is no corresponding record for this reference.
- 63Zimmerly, S.; Hausner, G.; Wu, X. C. Phylogenetic relationships among group II intron ORFs. Nucleic Acids Res. 2001, 29, 1238– 1250, DOI: 10.1093/nar/29.5.1238Google Scholar63https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXhvFGgsbk%253D&md5=a70b635e53fde84e69afc1f5b25a7d4fPhylogenetic relationships among group II intron ORFsZimmerly, Steven; Hausner, Georg; Wu, Xu-ChuNucleic Acids Research (2001), 29 (5), 1238-1250CODEN: NARHAD; ISSN:0305-1048. (Oxford University Press)Group II introns are widely believed to have been ancestors of spliceosomal introns, yet little is known about their own evolutionary history. In order to address the evolution of mobile group II introns, the authors have compiled 71 open reading frames (ORFs) related to group II intron reverse transcriptases and subjected their derived amino acid sequences to phylogenetic anal. The phylogenetic tree was rooted with reverse transcriptases (RTs) of non-long terminal repeat retroelements, and the inferred phylogeny reveals two major clusters which the authors term the mitochondrial and chloroplast-like lineages. Bacterial ORFs are mainly positioned at the bases of the two lineages but with weak bootstrap support. The data give an overview of an apparently high degree of horizontal transfer of group II intron ORFs, mostly among related organisms but also between organelles and. bacteria. The Zn domain (nuclease) and YADD motif (RT active site) were lost multiple times during evolution. Differences in domain structures suggest that the oldest ORFs were concise, while the ORF in the mitochondrial lineage subsequently expanded in three locations. The data are consistent with a bacterial origin for mobile group II introns.
- 64More, S. J.; Bampidis, V.; Benford, D.; Bragard, C.; Halldorsson, T. I.; Hernández-Jerez, A.; Susanne, H. B.; Koutsoumanis, K.; Machera, K.; Naegeli, H.; Nielsen, S. S.; Schlatter, J.; Schrenk, D.; Silano, V.; Turck, D.; Younes, M.; Glandorf, B.; Herman, L.; Tebbe, C.; Vlak, J.; Aguilera, J.; Schoonjans, R.; Cocconcelli, P. S. Evaluation of existing guidelines for their adequacy for the microbial characterisation and environmental risk assessment of microorganisms obtained through synthetic biology. EFSA J. 2020, 18, 1– 50Google ScholarThere is no corresponding record for this reference.
- 65Horton, R. M.; Hunt, H. D.; Ho, S. N.; Pullen, J. K.; Pease, L. R. Engineering hybrid genes without the use of restriction enzymes: gene splicing by overlap extension. Gene 1989, 77, 61– 68, DOI: 10.1016/0378-1119(89)90359-4Google Scholar65https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXktVais7k%253D&md5=40edd5de9a603f20bb901e78a3e3588dEngineering hybrid genes without the use of restriction enzymes: gene splicing by overlap extensionHorton, Robert M.; Hunt, Henry D.; Ho, Steffan N.; Pullen, Jeffrey K.; Pease, Larry R.Gene (1989), 77 (1), 61-8CODEN: GENED6; ISSN:0378-1119.Gene splicing by overlap extension is a new approach for recombining DNA mols. at precise junctions irresp. of nucleotide sequences at the recombination site and without the use of restriction endonucleases or ligase. Fragments from the genes that are to be recombined are generated in sep. polymerase chain reactions (PCRs). The primers are designed so that the ends of the products contain complementary sequences. When these PCR products are mixed, denatured, and reannealed, the strands having the matching sequences at their 3' ends overlap and act as primers for each other. Extension of this overlap by DNA polymerase produces a mol. in which the original sequences are spliced together. This technique is used to construct a gene encoding a mosaic fusion protein comprised of parts of two different mouse class-I major histocompatibility genes. This simple and widely applicable approach has significant advantages over std. recombinant DNA techniques.
- 66Gibson, D. G.; Young, L.; Chuang, R. Y.; Venter, J. C.; Hutchison, C. A.; Smith, H. O. Enzymatic assembly of DNA molecules up to several hundred kilobases. Nat. Methods 2009, 6, 343– 345, DOI: 10.1038/nmeth.1318Google Scholar66https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXksVemsbw%253D&md5=46284924c7d73c47cfb490983338e480Enzymatic assembly of DNA molecules up to several hundred kilobasesGibson, Daniel G.; Young, Lei; Chuang, Ray-Yuan; Venter, J. Craig; Hutchison, Clyde A.; Smith, Hamilton O.Nature Methods (2009), 6 (5), 343-345CODEN: NMAEA3; ISSN:1548-7091. (Nature Publishing Group)The authors describe an isothermal, single-reaction method for assembling multiple overlapping DNA mols. by the concerted action of a 5' exonuclease, a DNA polymerase and a DNA ligase. First they recessed DNA fragments, yielding single-stranded DNA overhangs that specifically annealed, and then covalently joined them. This assembly method can be used to seamlessly construct synthetic and natural genes, genetic pathways and entire genomes, and could be a useful mol. engineering tool.
- 67Gibson, D. G.; Glass, J. I.; Lartigue, C.; Noskov, V. N.; Chuang, R. Y.; Algire, M. A.; Benders, G. A.; Montague, M. G.; Ma, L.; Moodie, M. M.; Merryman, C.; Vashee, S.; Krishnakumar, R.; Assad-Garcia, N.; Andrews-Pfannkoch, C.; Denisova, E. A.; Young, L.; Qi, Z. N.; Segall-Shapiro, T. H.; Calvey, C. H.; Parmar, P. P.; Hutchison, C. A.; Smith, H. O.; Venter, J. C. Creation of a bacterial cell controlled by a chemically synthesized genome. Science 2010, 329, 52– 56, DOI: 10.1126/science.1190719Google Scholar67https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXotVeqsLg%253D&md5=f978c98f3e5c3c9b14bd4c7d8a1eecc7Creation of a bacterial cell controlled by a chemically synthesized genomeGibson, Daniel G.; Glass, John I.; Lartigue, Carole; Noskov, Vladimir N.; Chuang, Ray-Yuan; Algire, Mikkel A.; Benders, Gwynedd A.; Montague, Michael G.; Ma, Li; Moodie, Monzia M.; Merryman, Chuck; Vashee, Sanjay; Krishnakumar, Radha; Assad-Garcia, Nacyra; Andrews-Pfannkoch, Cynthia; Denisova, Evgeniya A.; Young, Lei; Qi, Zhi-Qing; Segall-Shapiro, Thomas H.; Calvey, Christopher H.; Parmar, Prashanth P.; Hutchison, Clyde A., III; Smith, Hamilton O.; Venter, J. CraigScience (Washington, DC, United States) (2010), 329 (5987), 52-56CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)We report the design, synthesis, and assembly of the 1.08-mega-base pair Mycoplasma mycoides JCVI-syn1.0 genome starting from digitized genome sequence information and its transplantation into a M. capricolum recipient cell to create new M. mycoides cells that are controlled only by the synthetic chromosome. The only DNA in the cells is the designed synthetic DNA sequence, including "watermark" sequences and other designed gene deletions and polymorphisms, and mutations acquired during the building process. The new cells have expected phenotypic properties and are capable of continuous self-replication. The complete, annotated synthetic genome sequence is deposited in GenBank/EMBL/DDBJ with accession no. CP002027.
Cited By
This article is cited by 4 publications.
- Lennart Schada von Borzyskowski, Helena Schulz-Mirbach, Mauricio Troncoso Castellanos, Francesca Severi, Paul A. Gómez-Coronado, Nicole Paczia, Timo Glatter, Arren Bar-Even, Steffen N. Lindner, Tobias J. Erb. Implementation of the β-hydroxyaspartate cycle increases growth performance of Pseudomonas putida on the PET monomer ethylene glycol. Metabolic Engineering 2023, 76 , 97-109. https://doi.org/10.1016/j.ymben.2023.01.011
- Esteban Martínez-García, Sofía Fraile, Elena Algar, Tomás Aparicio, Elena Velázquez, Belén Calles, Huseyin Tas, Blas Blázquez, Bruno Martín, Clara Prieto, Lucas Sánchez-Sampedro, Morten H H Nørholm, Daniel C Volke, Nicolas T Wirth, Pavel Dvořák, Lorea Alejaldre, Lewis Grozinger, Matthew Crowther, Angel Goñi-Moreno, Pablo I Nikel, Juan Nogales, Víctor de Lorenzo. SEVA 4.0: an update of the Standard European Vector Architecture database for advanced analysis and programming of bacterial phenotypes. Nucleic Acids Research 2023, 51
(D1)
, D1558-D1567. https://doi.org/10.1093/nar/gkac1059
- Jonathan Tellechea-Luzardo, Leanne Hobbs, Elena Velázquez, Lenka Pelechova, Simon Woods, Víctor de Lorenzo, Natalio Krasnogor. Versioning biological cells for trustworthy cell engineering. Nature Communications 2022, 13
(1)
https://doi.org/10.1038/s41467-022-28350-4
- Elena Velázquez, Yamal Al‐Ramahi, Víctor de Lorenzo. CRISPR/Cas9‐enhanced Targetron Insertion for Delivery of Heterologous Sequences into the Genome of Gram‐Negative Bacteria. Current Protocols 2022, 2
(9)
https://doi.org/10.1002/cpz1.532
Abstract
Figure 1
Figure 1. Diagram of targetrons and CRISPR/Cas9-mediated counterselection of insertions. The figure shows the technology developed in this article. Plasmid pSEVA-GIIi expresses the Ll.LtrB group II intron (empty or with cargo sequences cloned in the MluI site) and its IEP (LtrA) in the same transcriptional unit from the upstream promoter. The transcriptional unit is led by a retargeting region at 5′ (including exon 1 and the 5′ sequence of the Ll.LtrB intron), where three short sequences retrieved from the target gene (IBS, EBS1d, and EBS2) were engineered at given sites of the predicted transcript to secure its proper folding and retargeting (location of diagnostic primers is indicated). After transcription, the intronic RNA folds into a very conserved secondary structure and associates with LtrA to perform the splicing process from the exons. A lariat RNA is generated that remains attached to LtrA, forming a ribonucleoprotein (RNP) complex. This RNP scans DNA molecules until it finds the target site for retrohoming. Reverse splicing links covalently the intronic RNA to the sense strand of the DNA molecule, and the endonuclease (En) domain present in LtrA cleaves the antisense strand. Afterward, the retrotranscriptase (RT) domain of LtrA reverse transcribes the intronic RNA into DNA. The complete integration and synthesis of the cDNA is driven by repair mechanisms without the involvement of recombination. The incorporation of Cas9 complexes with gRNAs that recognize the insertion locus of Ll.LtrB causes the elimination of nonedited cells, as only those which incorporated the group II intron at the correct locus can survive (counterselection). IBS1: intron-binding site 1; EBS1d: exon-binding site 1-δ; EBS2: exon-binding site 2.
Figure 2
Figure 2. SEVA plasmids encoding the Ll.LtrB group II intron and T7 RNAP work in E. coli BL21DE3 and P. putida KT2440. (A) Delivery of the Ll.LtrB intron from plasmid pSEVA421-GIIi (Km) in E. coli BL21DE3. The Ll.LtrB intron was retargeted to insert in the antisense orientation into the locus 1063 of the lacZ gene so that insertions would disrupt this gene, giving rise to white colonies in the presence of X-gal. Since a RAM is placed inside Ll.LtrB, kanamycin resistance was used as a way to select for intron insertion mutants (plates to the right). (B) Graph shows the number of KmR CFU normalized to 109 viable cells and classified according to the displayed phenotype in the presence of X-gal (blue colonies: lacZ+, blue bars; white colonies: lacZ– (disrupted), white hatched bars) and, also, according to the presence or absence of IPTG induction. (C) Representative colony PCR reactions to determine the correct insertion of Ll.LtrB inside the lacZ gene. Only if Ll.LtrB retrohomes, a PCR amplicon of 720 bp is generated. Blue numbers correspond to blue colonies, and black numbers correspond to white colonies used as the template material for each reaction. (D–F) Delivery of Ll.LtrB from plasmid pSEVA421-GIIi-pyrF and with the help of pSEVA131-T7RNAP in P. putida KT2440. (D) Bar plot showing the frequency of 5FOAR CFU normalized to 109 viable cells after the insertion assay. CFU are classified according to the addition or absence of IPTG during the incubation period. (E) Genuine efficiency of the insertion of Ll.LtrB. The proportion of uracil auxotrophs detected from the 5FOAR population was used as a ratio to determine the abundance of Ll.LtrB insertions in the population. (F) 5FOA counterselection was used to isolate insertion mutants that were not able to grow without uracil supplemented to plates. Colonies resistant to 5FOA but that were able to grow without uracil were used as negative controls of insertion. Two different PCR reactions are shown: (top gel) one primer annealed inside Ll.LtrB and the second annealed in the pyrF gene so that an amplicon could be only generated after intron insertion. (Bottom gel) two primers flanking the insertion locus were used so that two amplicons could be generated. The smallest fragment (380 bp) corresponds to the WT sequence, and the biggest fragment (∼1500 bp) corresponds to the insertion. The same colonies were tested in both PCR reactions. All bar graphs show the mean values (bars), standard deviation (lines), and single values (dots) of two biological replicates. WT: wild-type; Mut: insertion mutant; + control: reaction with a colony with successful insertion from a previous experiment used as a template; H2O: control PCR with no template material; UraR: colonies growing in media without uracil (no auxotrophs); UraS: colonies not growing in media without uracil (auxotrophs).
Figure 3
Figure 3. Performance of pSEVA2311-GIIi-pyrF in P. putida KT2440 and its ΔrecA derivative strain. (A) pSEVA2311-GIIi-pyrF works in P. putida KT2440 to deliver the Ll.LtrB intron into the pyrF gene through 5FOA counterselection. Left panel: Bar plot showing the frequency of 5FOAR CFUs normalized to 109 viable cells of wild-type P. putida after the insertion assay resulting from induction with various cyclohexanone concentrations (0, 0.5, 1, and 5 mM). Right panel: Genuine efficiency of insertion of Ll.LtrB. The proportion of uracil auxotrophs detected from the 5FOAR population was used as a reference to determine the abundance of Ll.LtrB insertions in each population. (B) Colony PCR reaction that used primers flanking the insertion locus was used to determine Ll.LtrB retrohoming in each concentration of cyclohexanone tested (from 0 mM in the left part of gel to 5 mM in the right part of gel). The smallest fragment (380 bp) corresponds to the wild-type sequence, while the biggest one (1500 bp) corresponds to the intron insertion in the correct location. (C, D) Functioning of pSEVA2311-GIIi-pyrF in P. putida KT2440 ΔrecA. The same experiments and analyses were done to determine the frequencies and correctness of intron insertion in the recombination-deficient strain. The gel at the bottom of (D) shows a control PCR reaction to verify recA minus genotype of the tested cells. The frequency of 5FOAR CFUs and the efficiency of insertion of Ll.LtrB in P. putida ΔrecA were determined as before. All bar graphs show the mean values (bars), standard deviation (lines), and single values (dots) of at least three biological replicates. WT, wild-type; Mut, insertion mutant; KT2440, parental strain; + control: PCR of DNA from an intron-inserted colony (from a previous experiment) used as the template; H2O: control PCR with no template material.
Figure 4
Figure 4. Assessing the size-restriction of intron-mediated delivery with CRISPR/Cas9 counterselection using luxC fragments as a cargo. (A) Schematic of the intron library generated with increasing fragment length as a cargo (from 150 up to 1050 bp) using as template the first gene of the luxCDABEG operon, luxC. The Ll.LtrB intron in pSEVA6511-GIIi (LuxN) is retargeted to insert between the nucleotides 165 and 166 of the pyrF ORF in the antisense orientation. Spacer pyrF1 recognizes the region after the insertion site (part of the recognition site is shown inside a red box). The complementary nucleotides to the PAM (5′GGG-3′) are highlighted in red and are disrupted upon intron insertion. (B) Ll.LtrB-mediated delivery of luxC fragments in P. putida KT2440 WT with CRISPR/Cas9 counterselection. Colony PCR reactions showing amplifications from colonies with Ll.LtrB::LuxØ and Ll.LtrB::Lux1 (top gel) and corresponding PCR reactions verifying the recA-plus phenotype (bottom gel). WT amplification for the recA gene is 2 kb long. (C) Ll.LtrB-mediated delivery of luxC fragments in P. putida KT2440 ΔrecA. Colony PCR reactions showing amplifications from colonies with Ll.LtrB::LuxØ to Ll.LtrB::Lux4 (top gel) and corresponding PCR reaction verifying the recA—genotype (bottom gel). Deletion of the recA gene gives an amplification of 1 kb. WT: wild-type, Mut: insertion mutant, LuxN: cargos including from LuxØ to Lux7, Ø: Ll.LtrB with no cargo, 1: Ll.LtrB with Lux1 as the cargo, 2: Ll.LtrB with Lux2 as the cargo; 3: Ll.LtrB with Lux3 as the cargo, and 4: Ll.LtrB with Lux4 as the cargo. P. putida KT2440 ΔrecA colonies with no inserted Ll.LtrB used as the negative control.
Figure 5
Figure 5. Intron insertion frequencies with CRISPR-Cas9 counterselection in wild-type and ΔrecAP. putida KT2440 confirmed through PCR. (A) Intron insertion frequency of each cargo inserted in the genome of WT P. putida KT2440 by Ll.LtrB using 5FOA; no counterselection (control plasmid pSEVA231-CRISPR) or CRISPR/Cas9-mediated counterselection (pSEVA231-C-pyrF1). Insertion of a cargo larger than Lux1 was not detected. (B) Similarly, for P. putida KT2440 ΔrecA, insertion of a cargo larger than Lux4 was not detected. (C) Combined efficiency of targetrons and CRISPR/Cas9 counterselection. The numbers of SmR KmR CFU obtained after transforming either pSEVA231-CRISPR (control) or pSEVA231-C-pyrF1 were normalized to 109 cells. The pSEVA231-CRISPR condition was set to 100%, and the percentage of cells with pSEVA231-C-pyrF1 was calculated accordingly. Finally, the ratio of positive Ll.LtrB insertions detected in the pSEVA231-C-pyrF1 condition was multiplied individually in each replicate. The mean (bars), single values (dots), and standard deviation (lines) of two or three replicates are shown. Ø: Ll.LtrB with no cargo; 1: Ll.LtrB with Lux1 as the cargo; 2: Ll.LtrB with Lux2 as the cargo; 3: Ll.LtrB with Lux3 as the cargo; and 4: Ll.LtrB with Lux4 as the cargo.
Figure 6
Figure 6. Application of the Ll.LtrB group II intron for delivery of specific genetic barcodes to the genome of P. putida KT2440. Organization of pSEVA6511-GIIi (B3) variants is shown with the intron retargeted toward Locus 1 (x = 37s) or Locus 2 (x = 95a). Selection of the insertion loci for Ll.LtrB::B3 is in the vicinity of the Tn7-insertion site (black triangle). Two different insertion points (gray triangles) were chosen for the insertion list generated on the Clostron website, (24) and Ll.LtrB::B3 was retargeted to both sites accordingly. The recognition site in Locus 1 (green) is located in the sense strand, while Locus 2 (orange) is present in the antisense strand of the P. putida genome. Ll.LtrB::B3 insertion would generate two different genotypes depending on the locus being targeted in each case.
Figure 7
Figure 7. Delivery of Ll.LtrB::B3 containing a barcode in the P. putida KT2440 genome. A first pool PCR was set to detect successful Ll.LtrB::B3 insertions in either Locus 1 (green) or Locus 2 (orange). The top gel shows the amplification found with a pool PCR using primers flanking the insertion at Locus 2. The bottom gel shows the second PCR of individual colonies from the corresponding pool to find the barcoded clone. In this case, a primer annealing inside the barcode (pbarcode universal) and another annealing inside the PP5408 gene were used. WT: wild-type, Locus 1: 36,37s insertion site, Locus 2: 94,95a insertion site. PP5408 (green gene), glmS (orange gene).
References
ARTICLE SECTIONSThis article references 67 other publications.
- 1Nikel, P. I.; de Lorenzo, V. Pseudomonas putida as a functional chassis for industrial biocatalysis: From native biochemistry to trans-metabolism. Metab. Eng. 2018, 50, 142– 155, DOI: 10.1016/j.ymben.2018.05.005Google Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXpvVSlt7k%253D&md5=fac6e8937d2e4f3d0c26bcde01e6777cPseudomonas putida as a functional chassis for industrial biocatalysis: From native biochemistry to trans-metabolismNikel, Pablo I.; de Lorenzo, VictorMetabolic Engineering (2018), 50 (), 142-155CODEN: MEENFM; ISSN:1096-7176. (Elsevier B.V.)The itinerary followed by Pseudomonas putida from being a soil-dweller and plant colonizer bacterium to become a flexible and engineer-able platform for metabolic engineering stems from its natural lifestyle, which is adapted to harsh environmental conditions and all sorts of physicochem. stresses. Over the years, these properties have been capitalized biotechnol. owing to the expanding wealth of genetic tools designed for deep-editing the P. putida genome. A suite of dedicated vectors inspired in the core tenets of synthetic biol. have enabled to suppress many of the naturally-occurring undesirable traits native to this species while enhancing its many appealing properties, and also to import catalytic activities and attributes from other biol. systems. Much of the biotechnol. interest on P. putida stems from the distinct architecture of its central carbon metab. The native biochem. is naturally geared to generate reductive currency [i.e., NAD(P)H] that makes this bacterium a phenomenal host for redox-intensive reactions. In some cases, genetic editing of the indigenous biochem. network of P. putida (cis-metab.) has sufficed to obtain target compds. of industrial interest. Yet, the main value and promise of this species (in particular, strain KT2440) resides not only in its capacity to host heterologous pathways from other microorganisms, but also altogether artificial routes (trans-metab.) for making complex, new-to-Nature mols. A no. of examples are presented for substantiating the worth of P. putida as one of the favorite workhorses for sustainable manufg. of fine and bulk chems. in the current times of the 4th Industrial Revolution. The potential of P. putida to extend its rich native biochem. beyond existing boundaries is discussed and research bottlenecks to this end are also identified. These aspects include not just the innovative genetic design of new strains but also the incorporation of novel chem. elements into the extant biochem., as well as genomic stability and scaling-up issues.
- 2Nikel, P. I.; Martínez-García, E.; de Lorenzo, V. Biotechnological domestication of pseudomonads using synthetic biology. Nat. Rev. Microbiol. 2014, 12, 368– 379, DOI: 10.1038/nrmicro3253Google Scholar2https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXmtlahtL8%253D&md5=d2cc342c2b17f5631572b664f5513acbBiotechnological domestication of pseudomonads using synthetic biologyNikel, Pablo I.; Martinez-Garcia, Esteban; de Lorenzo, VictorNature Reviews Microbiology (2014), 12 (5), 368-379CODEN: NRMACK; ISSN:1740-1526. (Nature Publishing Group)A review. Much of contemporary synthetic biol. research relies on the use of bacterial chassis for plugging-in and plugging-out genetic circuits and new-to-nature functionalities. However, the microorganisms that are the easiest to manipulate in the lab. are often suboptimal for downstream industrial applications, which can involve physicochem. stress and harsh operating conditions. In this Review, we advocate the use of environmental Pseudomonas strains as model organisms that are pre-endowed with the metabolic, physiol. and stress-endurance traits that are demanded by current and future synthetic biol. and biotechnol. needs.
- 3Kampers, L. F. C.; Volkers, R. J. M.; Martins Dos Santos, V. A. P. Pseudomonas putida KT2440 is HV1 certified, not GRAS. Microb. Biotechnol. 2019, 12, 845– 848, DOI: 10.1111/1751-7915.13443Google Scholar3https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BB3M3msleqsA%253D%253D&md5=01a803206ea94a4286be8ab8a039549cPseudomonas putida KT2440 is HV1 certified, not GRASKampers Linde F C; Volkers Rita J M; Martins Dos Santos Vitor A P; Martins Dos Santos Vitor A PMicrobial biotechnology (2019), 12 (5), 845-848 ISSN:.Pseudomonas putida is rapidly becoming a workhorse for industrial production due to its metabolic versatility, genetic accessibility and stress-resistance properties. The P. putida strain KT2440 is often described as Generally Regarded as Safe, or GRAS, indicating the strain is safe to use as food additive. This description is incorrect. P. putida KT2440 is classified by the FDA as HV1 certified, indicating it is safe to use in a P1 or ML1 environment.
- 4Kim, J.; Park, W. Oxidative stress response in Pseudomonas putida. Appl. Microbiol. Biotechnol. 2014, 98, 6933– 6946, DOI: 10.1007/s00253-014-5883-4Google Scholar4https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhtVamtLbN&md5=de993728cab8409fdd59ed5c00fcb724Oxidative stress response in Pseudomonas putidaKim, Jisun; Park, WoojunApplied Microbiology and Biotechnology (2014), 98 (16), 6933-6946CODEN: AMBIDG; ISSN:0175-7598. (Springer)A review. Pseudomonas putida is widely distributed in nature and is capable of degrading various org. compds. due to its high metabolic versatility. The survival capacity of P. putida stems from its frequent exposure to various endogenous and exogenous oxidative stresses. Oxidative stress is an unavoidable consequence of interactions with various reactive oxygen species (ROS)-inducing agents existing in various niches. ROS could facilitate the evolution of bacteria by mutating genomes. Aerobic bacteria maintain defense mechanisms against oxidative stress throughout their evolution. To overcome the detrimental effects of oxidative stress, P. putida has developed defensive cellular systems involving induction of stress-sensing proteins and detoxification enzymes as well as regulation of oxidative stress response networks. Genetic responses to oxidative stress in P. putida differ markedly from those obsd. in Escherichia coli and Salmonella spp. Two major redox-sensing transcriptional regulators, SoxR and OxyR, are present and functional in the genome of P. putida. However, the novel regulators FinR and HexR control many genes belonging to the E. coli SoxR regulon. Oxidative stress can be generated by exposure to antibiotics, and iron homeostasis in P. putida is crucial for bacterial cell survival during treatment with antibiotics. This review highlights and summarizes current knowledge of oxidative stress in P. putida, as a model soil bacterium, together with recent studies from mol. genetics perspectives.
- 5Nikel, P. I.; Fuhrer, T.; Chavarría, M.; Sánchez-Pascuala, A.; Sauer, U.; de Lorenzo, V. Reconfiguration of metabolic fluxes in Pseudomonas putida as a response to sub-lethal oxidative stress. ISME J. 2021, 15, 1751– 1766, DOI: 10.1038/s41396-020-00884-9Google Scholar5https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXitVSisLc%253D&md5=31d1780e11c16b4bbdbf09c008f4324dReconfiguration of metabolic fluxes in Pseudomonas putida as a response to sub-lethal oxidative stressNikel, Pablo I.; Fuhrer, Tobias; Chavarria, Max; Sanchez-Pascuala, Alberto; Sauer, Uwe; de Lorenzo, VictorISME Journal (2021), 15 (6), 1751-1766CODEN: IJSOCF; ISSN:1751-7362. (Nature Portfolio)As a frequent inhabitant of sites polluted with toxic chems., the soil bacterium and plant-root colonizer Pseudomonas putida can tolerate high levels of endogenous and exogenous oxidative stress. Yet, the ultimate reason of such phenotypic property remains largely unknown. To shed light on this question, metabolic network-wide routes for NADPH generation-the metabolic currency that fuels redox-stress quenching mechanisms-were inspected when P. putida KT2440 was challenged with a sub-lethal H2O2 dose as a proxy of oxidative conditions. 13C-tracer expts., metabolomics, and flux anal., together with the assessment of physiol. parameters and measurement of enzymic activities, revealed a substantial flux reconfiguration in oxidative environments. In particular, periplasmic glucose processing was rerouted to cytoplasmic oxidn., and the cyclic operation of the pentose phosphate pathway led to significant NADPH-forming fluxes, exceeding biosynthetic demands by ∼50%. The resulting NADPH surplus, in turn, fueled the glutathione system for H2O2 redn. These properties not only account for the tolerance of P. putida to environmental insults-some of which end up in the formation of reactive oxygen species-but they also highlight the value of this bacterial host as a platform for environmental bioremediation and metabolic engineering.
- 6Liao, J. C.; Mi, L.; Pontrelli, S.; Luo, S. Fuelling the future: Microbial engineering for the production of sustainable biofuels. Nat. Rev. Microbiol. 2016, 14, 288– 304, DOI: 10.1038/nrmicro.2016.32Google Scholar6https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28XltVaqsbk%253D&md5=fcd67bab889558ff6f918a5d70a9d6ecFuelling the future: microbial engineering for the production of sustainable biofuelsLiao, James C.; Mi, Luo; Pontrelli, Sammy; Luo, ShanshanNature Reviews Microbiology (2016), 14 (5), 288-304CODEN: NRMACK; ISSN:1740-1526. (Nature Publishing Group)Global climate change linked to the accumulation of greenhouse gases has caused concerns regarding the use of fossil fuels as the major energy source. To mitigate climate change while keeping energy supply sustainable, one soln. is to rely on the ability of microorganisms to use renewable resources for biofuel synthesis. In this Review, we discuss how microorganisms can be explored for the prodn. of next-generation biofuels, based on the ability of bacteria and fungi to use lignocellulose; through direct CO2 conversion by microalgae; using lithoautotrophs driven by solar electricity; or through the capacity of microorganisms to use methane generated from landfill. Furthermore, we discuss how to direct these substrates to the biosynthetic pathways of various fuel compds. and how to optimize biofuel prodn. by engineering fuel pathways and central metab.
- 7Jiménez, J. I.; Miñambres, B.; García, J. L.; Díaz, E. Genomic analysis of the aromatic catabolic pathways from Pseudomonas putida KT2440. Environ. Microbiol. 2002, 4, 824– 841, DOI: 10.1046/j.1462-2920.2002.00370.xGoogle Scholar7https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXkvVGrtg%253D%253D&md5=1922f12af80e022107400b58fa47f419Genomic analysis of the aromatic catabolic pathways from Pseudomonas putida KT2440Jimenez, Jose Ignacio; Minambres, Baltasar; Garcia, Jose Luis; Diaz, EduardoEnvironmental Microbiology (2002), 4 (12), 824-841CODEN: ENMIFM; ISSN:1462-2912. (Blackwell Science Ltd.)Anal. of the catabolic potential of Pseudomonas putida KT2440 against a wide range of natural arom. compds. and sequence comparisons with the entire genome of this microorganism predicted the existence of at least four main pathways for the catabolism of central arom. intermediates, i.e., the protocatechuate (pca genes) and catechol (cat genes) branches of the β-ketoadipate pathway, the homogentisate pathway (hmg/fah/mai genes) and the phenylacetate pathway (pha genes). Two addnl. gene clusters that might be involved in the catabolism of N-heterocyclic arom. compds. (nic cluster) and in a central meta-cleavage pathway (pcm genes) were also identified. Furthermore, the genes encoding the peripheral pathways for the catabolism of p-hydroxybenzoate (pob), benzoate (ben), quinate (qui), phenylpropenoid compds. (fcs, ech, vdh, cal, van, acd and acs), phenylalanine and tyrosine (phh, hpd) and n-phenylalkanoic acids (fad) were mapped in the chromosome of P. putida KT2440. Although a repetitive extragenic palindromic (REP) element is usually assocd. with the gene clusters, a supraoperonic clustering of catabolic genes that channel different arom. compds. into a common central pathway (catabolic island) was not obsd. in P. putida KT2440. The global view on the mineralization of arom. compds. by P. putida KT2440 will facilitate the rational manipulation of this strain for improving biodegrdn./biotransformation processes, and reveals this bacterium as a useful model system for studying biochem., genetic, evolutionary and ecol. aspects of the catabolism of arom. compds.
- 8Ramos, J. L.; Duque, E.; Huertas, M. J.; Haidour, A. Isolation and expansion of the catabolic potential of a Pseudomonas putida strain able to grow in the presence of high concentrations of aromatic hydrocarbons. J. Bacteriol. 1995, 177, 3911– 3916, DOI: 10.1128/jb.177.14.3911-3916.1995Google Scholar8https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXmvV2gsb0%253D&md5=92a84eea5b4f2a8abba19a91b534c80bIsolation and expansion of the catabolic potential of a Pseudomonas putida strain able to grow in the presence of high concentrations of aromatic hydrocarbonsRamos, Juan L.; Duque, Estrella; Huertas, Maria-Jose; Haidour, AliJournal of Bacteriology (1995), 177 (14), 3911-16CODEN: JOBAAY; ISSN:0021-9193. (American Society for Microbiology)Pseudomonas putida DOT-T1 was isolated after enrichment on minimal medium with 1% (vol/vol) toluene as the sole C source. The strain was able to grow in the presence of 90% (vol/vol) toluene and was tolerant to org. solvents whose log Pow (octanol/water partition coeff.) was higher than 2.3. Solvent tolerance was inducible, as bacteria grown in the absence of toluene required an adaptation period before growth restarted. Mg2+ ions in the culture medium improved solvent tolerance. Electron micrographs showed that cells growing on high concns. of toluene exhibited a wider periplasmic space than cells growing in the absence of toluene and preserved the outer membrane integrity. Polarog. studies and the accumulation of pathway intermediates showed that the strain used the toluene-4-monooxygenase pathway to catabolize toluene. Although the strain also thrived in high concns. of m- and p-xylene, these hydrocarbons could not be used as the sole C source for growth. The catabolic potential of the isolate was expanded to include m- and p-xylene and related hydrocarbons by transfer of the TOL plasmid pWW0-Km.
- 9Martínez, I.; Mohamed, M. E. S.; Rozas, D.; García, J. L.; Díaz, E. Engineering synthetic bacterial consortia for enhanced desulfurization and revalorization of oil sulfur compounds. Metab. Eng. 2016, 35, 46– 54, DOI: 10.1016/j.ymben.2016.01.005Google Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC28Xhtlalu70%253D&md5=6021bcdce6cb854cabf99a30150e080aEngineering synthetic bacterial consortia for enhanced desulfurization and revalorization of oil sulfur compoundsMartinez, Igor; Mohamed, Magdy El-Said; Rozas, Daniel; Garcia, Jose Luis; Diaz, EduardoMetabolic Engineering (2016), 35 (), 46-54CODEN: MEENFM; ISSN:1096-7176. (Elsevier B. V.)The 4S pathway is the most studied bioprocess for the removal of the recalcitrant sulfur of arom. heterocycles present in fuels. It consists of three sequential functional units, encoded by the dszABCD genes, through which the model compd. dibenzothiophene (DBT) is transformed into the sulfur-free 2-hydroxybiphenyl (2HBP) mol. In this work, a set of synthetic dsz cassettes were implanted in Pseudomonas putida KT2440, a model bacterial "chassis" for metabolic engineering studies. The complete dszB1A1C1-D1 cassette behaved as an attractive alternative - to the previously constructed recombinant dsz cassettes - for the conversion of DBT into 2HBP. Refactoring the 4S pathway by the use of synthetic dsz modules encoding individual 4S pathway reactions revealed unanticipated traits, e.g., the 4S intermediate 2HBP-sulfinate (HBPS) behaves as an inhibitor of the Dsz monooxygenases, and once secreted from the cells it cannot be further taken up. That issue should be addressed for the rational design of more efficient biocatalysts for DBT bioconversions. In this sense, the construction of synthetic bacterial consortia to compartmentalize the 4S pathway into different cell factories for individual optimization was shown to enhance the conversion of DBT into 2HBP, overcome the inhibition of the Dsz enzymes by the 4S intermediates, and enable efficient prodn. of unattainable high added value intermediates, e.g., HBPS, that are difficult to obtain using the current monocultures.
- 10Martínez-García, E.; Nikel, P. I.; Aparicio, T.; de Lorenzo, V. Pseudomonas 2.0: Genetic upgrading of P. putida KT2440 as an enhanced host for heterologous gene expression. Microb. Cell Fact. 2014, 13, 159 DOI: 10.1186/s12934-014-0159-3Google Scholar10https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXit1Oqt7g%253D&md5=dfebf084cef553ec5d6526f9fbbd504fPseudomonas 2.0: genetic upgrading of P. putida KT2440 as an enhanced host for heterologous gene expressionMartinez-Garcia, Esteban; Nikel, Pablo I.; Aparicio, Tomas; de Lorenzo, VictorMicrobial Cell Factories (2014), 13 (), 159/1-159/33CODEN: MCFICT; ISSN:1475-2859. (BioMed Central Ltd.)Background: Because of its adaptability to sites polluted with toxic chems., the model soil bacterium Pseudomonas putida is naturally endowed with a no. of metabolic and stress-endurance qualities which have considerable value for hosting energy-demanding and redox reactions thereof. The growing body of knowledge on P. putida strain KT2440 has been exploited for the rational design of a deriv. strain in which the genome has been heavily edited in order to construct a robust microbial cell factory. Results: Eleven non-adjacent genomic deletions, which span 300 genes (i.e., 4.3% of the entire P. putida KT2440 genome), were eliminated; thereby enhancing desirable traits and eliminating attributes which are detrimental in an expression host. Since ATP and NAD(P)H availability - as well as genetic instability, are generally considered to be major bottlenecks for the performance of platform strains, a suite of functions that drain high-energy phosphate from the cells and/or consume NAD(P)H were targeted in particular, the whole flagellar machinery. Four prophages, two transposons, and three components of DNA restriction modification systems were eliminated as well. The resulting strain (P. putida EM383) displayed growth properties (i.e., lag times, biomass yield, and specific growth rates) clearly superior to the precursor wild-type strain KT2440. Furthermore, it tolerated endogenous oxidative stress, acquired and replicated exogenous DNA, and survived better in stationary phase. The performance of a bi-cistronic GFP-LuxCDABE reporter system as a proxy of combined metabolic vitality, revealed that the deletions in P. putida strain EM383 brought about an increase of >50% in the overall physiol. vigour. Conclusion: The rationally modified P. putida strain allowed for the better functional expression of implanted genes by directly improving the metabolic currency that sustains the gene expression flow, instead of resorting to the classical genetic approaches (e.g., increasing the promoter strength in the DNA constructs of interest).
- 11Lieder, S.; Nikel, P. I.; de Lorenzo, V.; Takors, R. Genome reduction boosts heterologous gene expression in Pseudomonas putida. Microb. Cell Fact. 2015, 14, 23 DOI: 10.1186/s12934-015-0207-7Google Scholar11https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC2MjlvVeqtA%253D%253D&md5=1b3829a02ebcccfb6bd3f5dfc0299255Genome reduction boosts heterologous gene expression in Pseudomonas putidaLieder Sarah; Takors Ralf; Nikel Pablo I; de Lorenzo VictorMicrobial cell factories (2015), 14 (), 23 ISSN:.BACKGROUND: The implementation of novel platform organisms to be used as microbial cell factories in industrial applications is currently the subject of intense research. Ongoing efforts include the adoption of Pseudomonas putida KT2440 variants with a reduced genome as the functional chassis for biotechnological purposes. In these strains, dispensable functions removed include flagellar motility (1.1% of the genome) and a number of open reading frames expected to improve genotypic and phenotypic stability of the cells upon deletion (3.2% of the genome). RESULTS: In this study, two previously constructed multiple-deletion P. putida strains were systematically evaluated as microbial cell factories for heterologous protein production and compared to the parental bacterium (strain KT2440) with regards to several industrially-relevant physiological traits. Energetic parameters were quantified at different controlled growth rates in continuous cultivations and both strains had a higher adenosine triphosphate content, increased adenylate energy charges, and diminished maintenance demands than the wild-type strain. Under all the conditions tested the mutants also grew faster, had enhanced biomass yields and showed higher viability, and displayed increased plasmid stability than the parental strain. In addition to small-scale shaken-flask cultivations, the performance of the genome-streamlined strains was evaluated in larger scale bioreactor batch cultivations taking a step towards industrial growth conditions. When the production of the green fluorescent protein (used as a model heterologous protein) was assessed in these cultures, the mutants reached a recombinant protein yield with respect to biomass up to 40% higher than that of P. putida KT2440. CONCLUSIONS: The two streamlined-genome derivatives of P. putida KT2440 outcompeted the parental strain in every industrially-relevant trait assessed, particularly under the working conditions of a bioreactor. Our results demonstrate that these genome-streamlined bacteria are not only robust microbial cell factories on their own, but also a promising foundation for further biotechnological applications.
- 12Silva-Rocha, R.; Martínez-García, E.; Calles, B.; Chavarría, M.; Arce-Rodríguez, A.; De Las Heras, A.; Páez-Espino, A. D.; Durante-Rodríguez, G.; Kim, J.; Nikel, P. I.; Platero, R.; De Lorenzo, V. The Standard European Vector Architecture (SEVA): A coherent platform for the analysis and deployment of complex prokaryotic phenotypes. Nucleic Acids Res. 2013, 41, D666– D675, DOI: 10.1093/nar/gks1119Google Scholar12https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38XhvV2ktrzF&md5=3ddd60ae5099d577aa6b688813a5605bThe Standard European Vector Architecture (SEVA): a coherent platform for the analysis and deployment of complex prokaryotic phenotypesSilva-Rocha, Rafael; Martinez-Garcia, Esteban; Calles, Belen; Chavarria, Max; Arce-Rodriguez, Alejandro; de las Heras, Aitor; Paez-Espino, A. David; Durante-Rodriguez, Gonzalo; Kim, Juhyun; Nikel, Pablo I.; Platero, Raul; de Lorenzo, VictorNucleic Acids Research (2013), 41 (D1), D666-D675CODEN: NARHAD; ISSN:0305-1048. (Oxford University Press)The Std. European Vector Architecture' database (SEVA-DB, http://seva.cnb.csic.es) was conceived as a user-friendly, web-based resource and a material clone repository to assist in the choice of optimal plasmid vectors for de-constructing and re-constructing complex prokaryotic phenotypes. The SEVA-DB adopts simple design concepts that facilitate the swapping of functional modules and the extension of genome engineering options to microorganisms beyond typical lab. strains. Under the SEVA std., every DNA portion of the plasmid vectors is minimized, edited for flaws in their sequence and/or functionality, and endowed with phys. connectivity through three inter-segment insulators that are flanked by fixed, rare restriction sites. Such a scaffold enables the exchangeability of multiple origins of replication and diverse antibiotic selection markers to shape a frame for their further combination with a large variety of cargo modules that can be used for varied end-applications. The core collection of constructs that are available at the SEVA-DB has been produced as a starting point for the further expansion of the formatted vector platform. We argue that adoption of the SEVA format can become a shortcut to fill the phenomenal gap between the existing power of DNA synthesis and the actual engineering of predictable and efficacious bacteria.
- 13Martínez-García, E.; de Lorenzo, V. Engineering multiple genomic deletions in Gram-negative bacteria: Analysis of the multi-resistant antibiotic profile of Pseudomonas putida KT2440. Environ. Microbiol. 2011, 13, 2702– 2716, DOI: 10.1111/j.1462-2920.2011.02538.xGoogle Scholar13https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3MXhsVKhtbfP&md5=8a2204c9c029721125df30a7c2ac80e2Engineering multiple genomic deletions in gram-negative bacteria: analysis of the multi-resistant antibiotic profile of Pseudomonas putida KT2440Martinez-Garcia, Esteban; de Lorenzo, VictorEnvironmental Microbiology (2011), 13 (10), 2702-2716CODEN: ENMIFM; ISSN:1462-2912. (Wiley-Blackwell)The genome of the soil bacterium Pseudomonas putida strain KT2440 has been erased of various determinants of resistance to antibiotics encoded in its extant chromosome. To this end, the authors employed a coherent genetic platform that allowed the precise deletion of multiple genomic segments in a large variety of Gram-neg. bacteria including (but not limited to) P. putida. The method is based on the obligatory recombination between free-ended homologous DNA sequences that are released as linear fragments generated upon the cleavage of the chromosome with unique I-SceI sites, added to the segment of interest by the vector system. Despite the potential for a SOS response brought about by the appearance of double stranded DNA breaks during the process, fluctuation expts. revealed that the procedure did not increase mutation rates - perhaps due to the protection exerted by I-SceI bound to the otherwise naked DNA termini. With this tool in hand the authors made sequential deletions of genes mexC, mexE, ttgA and ampC in the genome of the target bacterium, orthologues of which are known to det. various degrees of antibiotic resistance in diverse microorganisms. Inspection of the corresponding phenotypes demonstrated that the efflux pump encoded by ttgA sufficed to endow P. putida with a high-level of tolerance to β-lactams, chloramphenicol and quinolones, but had little effect on, e.g. aminoglycosides. Anal. of the mutants revealed also a considerable diversity in the manifestation of the resistance phenotype within the population and suggested a degree of synergism between different pumps. The directed edition of the P. putida chromosome shown here not only enhances the amenability of this bacterium to deep genomic engineering, but also validates the corresponding approach for similar handlings of a large variety of Gram-neg. microorganisms.
- 14Martínez-García, E.; De Lorenzo, V. Transposon-based and plasmid-based genetic tools for editing genomes of Gram-negative bacteria. Methods Mol. Biol. 2012, 813, 267– 283Google Scholar14https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xhs1Sqs7jK&md5=d8dc7fef841f41065b1e0c236b3339d8Transposon-based and plasmid-based genetic tools for editing genomes of gram-negative bacteriaMartinez-Garcia, Esteban; de Lorenzo, VictorMethods in Molecular Biology (New York, NY, United States) (2012), 813 (Synthetic Gene Networks), 267-283CODEN: MMBIED; ISSN:1064-3745. (Springer)A good part of the contemporary synthetic biol. agenda aims at reprogramming microorganisms to enhance existing functions and/or perform new tasks. Moreover, the functioning of complex regulatory networks, or even a single gene, is revealed only when perturbations are entered in the corresponding dynamic systems and the outcome monitored. These endeavors rely on the availability of genetic tools to successfully modify a la carte the chromosome of target bacteria. Key aspects to this end include the removal of undesired genomic segments, systems for the prodn. of directed mutants and allelic replacements, random mutant libraries to discover new functions, and means to stably implant larger genetic networks into the genome of specific hosts. The list of gram-neg. species that are appealing for such genetic refactoring operations is growingly expanding. However, the repertoire of available mol. techniques to do so is very limited beyond Escherichia coli. In this chapter, utilization of novel tools is described (exemplified in two plasmids systems: pBAM1 and pEMG) tailored for facilitating chromosomal engineering procedures in a wide variety of gram-neg. microorganisms.
- 15Martínez-García, E.; Aparicio, T.; de Lorenzo, V.; Nikel, P. I. New transposon tools tailored for metabolic engineering of Gram-negative microbial cell factories. Front. Bioeng. Biotechnol. 2014, 2, 46 DOI: 10.3389/fbioe.2014.00046Google Scholar15https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC2M3nsVOrtw%253D%253D&md5=e535341c401daaac2882e3d09ba452b3New transposon tools tailored for metabolic engineering of gram-negative microbial cell factoriesMartinez-Garcia Esteban; Aparicio Tomas; de Lorenzo Victor; Nikel Pablo IFrontiers in bioengineering and biotechnology (2014), 2 (), 46 ISSN:2296-4185.Re-programming microorganisms to modify their existing functions and/or to bestow bacteria with entirely new-to-Nature tasks have largely relied so far on specialized molecular biology tools. Such endeavors are not only relevant in the burgeoning metabolic engineering arena but also instrumental to explore the functioning of complex regulatory networks from a fundamental point of view. A la carte modification of bacterial genomes thus calls for novel tools to make genetic manipulations easier. We propose the use of a series of new broad-host-range mini-Tn5-vectors, termed pBAMDs, for the delivery of gene(s) into the chromosome of Gram-negative bacteria and for generating saturated mutagenesis libraries in gene function studies. These delivery vectors endow the user with the possibility of easy cloning and subsequent insertion of functional cargoes with three different antibiotic-resistance markers (kanamycin, streptomycin, and gentamicin). After validating the pBAMD vectors in the environmental bacterium Pseudomonas putida KT2440, their use was also illustrated by inserting the entire poly(3-hydroxybutyrate) (PHB) synthesis pathway from Cupriavidus necator in the chromosome of a phosphotransacetylase mutant of Escherichia coli. PHB is a completely biodegradable polyester with a number of industrial applications that make it attractive as a potential replacement of oil-based plastics. The non-selective nature of chromosomal insertions of the biosynthetic genes was evidenced by a large landscape of PHB synthesis levels in independent clones. One clone was selected and further characterized as a microbial cell factory for PHB accumulation, and it achieved polymer accumulation levels comparable to those of a plasmid-bearing recombinant. Taken together, our results demonstrate that the new mini-Tn5-vectors can be used to confer interesting phenotypes in Gram-negative bacteria that would be very difficult to engineer through direct manipulation of the structural genes.
- 16Aparicio, T.; Nyerges, A.; Martínez-García, E.; de Lorenzo, V. High-Efficiency Multi-site Genomic Editing of Pseudomonas putida through Thermoinducible ssDNA Recombineering. iScience 2020, 23, 100946 DOI: 10.1016/j.isci.2020.100946Google Scholar16https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXmsl2mtLg%253D&md5=7b4bae6ce59f76035cfa133e40ac0be5High-Efficiency Multi-site Genomic Editing of Pseudomonas putida through Thermoinducible ssDNA RecombineeringAparicio, Tomas; Nyerges, Akos; Martinez-Garcia, Esteban; de Lorenzo, VictoriScience (2020), 23 (3), 100946CODEN: ISCICE; ISSN:2589-0042. (Elsevier B.V.)Application of single-stranded DNA recombineering for genome editing of species other than enterobacteria is limited by the efficiency of the recombinase and the action of endogenous mismatch repair (MMR) systems. In this work we have set up a genetic system for entering multiple changes in the chromosome of the biotechnol. relevant strain EM42 of Pseudomononas putida. To this end high-level heat-inducible co-transcription of the rec2 recombinase and P. putida's allele mutLPPE36K was designed under the control of the PL/cI857 system. Cycles of short thermal shifts followed by transformation with a suite of mutagenic oligos delivered different types of genomic changes at frequencies up to 10% per single modification. The same approach was instrumental to super-diversify short chromosomal portions for creating libraries of functional genomic segments-e.g., ribosomal-binding sites. These results enabled multiplexing of genome engineering of P. putida, as required for metabolic reprogramming of this important synthetic biol. chassis.
- 17Martin-Pascual, M.; Batianis, C.; Bruinsma, L.; Asin-Garcia, E.; Garcia-Morales, L.; Weusthuis, R. A.; van Kranenburg, R.; Martins Dos Santos, V. A. P. A navigation guide of synthetic biology tools for Pseudomonas putida. Biotechnol. Adv. 2021, 49, 107732 DOI: 10.1016/j.biotechadv.2021.107732Google Scholar17https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXhs1yrt7bL&md5=3012891782213362ce45205f20b9ddfeA navigation guide of synthetic biology tools for Pseudomonas putidaMartin-Pascual, Maria; Batianis, Christos; Bruinsma, Lyon; Asin-Garcia, Enrique; Garcia-Morales, Luis; Weusthuis, Ruud A.; van Kranenburg, Richard; Martins dos Santos, Vitor A. P.Biotechnology Advances (2021), 49 (), 107732CODEN: BIADDD; ISSN:0734-9750. (Elsevier Inc.)Pseudomonas putida is a microbial chassis of huge potential for industrial and environmental biotechnol., owing to its remarkable metabolic versatility and ability to sustain difficult redox reactions and operational stresses, among other attractive characteristics. A wealth of genetic and in silico tools have been developed to enable the unravelling of its physiol. and improvement of its performance. However, the rise of this microbe as a promising platform for biotechnol. applications has resulted in diversification of tools and methods rather than standardization and convergence. As a consequence, multiple tools for the same purpose have been generated, while most of them have not been embraced by the scientific community, which has led to compartmentalization and inefficient use of resources. Inspired by this and by the substantial increase in popularity of P. putida, we aim herein to bring together and assess all currently available (wet and dry) synthetic biol. tools specific for this microbe, focusing on the last 5 years. We provide information on the principles, functionality, advantages and limitations, with special focus on their use in metabolic engineering. Addnl., we compare the tool portfolio for P. putida with those for other bacterial chassis and discuss potential future directions for tool development. Therefore, this review is intended as a ref. guide for experts and new 'users' of this promising chassis.
- 18Mörl, M.; Niemer, I.; Schmelzer, C. New reactions catalyzed by a group II intron ribozyme with RNA and DNA substrates. Cell 1992, 70, 803– 810, DOI: 10.1016/0092-8674(92)90313-2Google Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADyaK38zos1Gquw%253D%253D&md5=0de88e6dcdb98d1f7c6c68cdac24d25eNew reactions catalyzed by a group II intron ribozyme with RNA and DNA substratesMorl M; Niemer I; Schmelzer CCell (1992), 70 (5), 803-10 ISSN:0092-8674.Here we describe three novel reactions of the self-splicing group II intron bI1 (the first intron of the COB gene of yeast mitochondria) demonstrating its catalytic versatility: reversal of the first step of the self-splicing reaction catalyzed by a linear form of the intron utilizing the energy of a phosphoanhydride bond for transesterification, ligation of a single-stranded DNA to an RNA, and cleavage of a single-stranded DNA substrate. These results have the following evolutionary implications: use of the alpha-beta bond of a terminal triphosphate for transesterification suggests that an RNA RNA replicase could use mononucleotide triphosphates as precursors, and cleavage of single-stranded DNA and DNA-RNA ligation suggests that excised group II introns might integrate directly into DNA without prior reverse transcription.
- 19Lazowska, J.; Meunier, B.; Macadre, C. Homing of a group II intron in yeast mitochondrial DNA is accompanied by unidirectional co-conversion of upstream-located markers. EMBO J. 1994, 13, 4963– 4972, DOI: 10.1002/j.1460-2075.1994.tb06823.xGoogle Scholar19https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXhvF2hsrY%253D&md5=c7313b7d97832fa83b5ed5c1ba63683bHoming of a group II intron in yeast mitochondrial DNA is accompanied by unidirectional co-conversion of upstream-located markersLazowska, Jaga; Meunier, Brigitte; Macadre, CatherineEMBO Journal (1994), 13 (20), 4963-72CODEN: EMJODG; ISSN:0261-4189. (Oxford University Press)Group II introns ai1 and ai2 of the Saccharomyces cerevisiae mitochondrial COX1 gene encode proteins having a dual function (maturase and reverse transcriptase) and are mobile genetic elements. By construction of adequate donor genomes, we demonstrate that each of them is self-sufficient and practises homing in the absence of homing-type endonucleases encoded by either group I introns or the ENS2 gene. Each of the S. cerevisiae group II self-mobile introns was tested for its ability to invade mitochondrial DNA (mtDNA) from two related Saccharomyces species. Surprisingly, only ai2 was obsd. to integrate into both genomes. The non-mobility of ai1 was clearly correlated with some polymorphic changes occurring in sequences flanking its insertion sites in the recipient mtDNAs. Importantly, studies of the behavior of these introns in interspecific crosses demonstrate that flanking marker co-conversion accompanying group II intron homing is unidirectional and efficient only in the 3' to 5' direction towards the upstream exon. Thus, the polar co-conversion and dependence of the splicing proficiency of the intron reported previously by us are hallmarks of group II intron homing, which significantly distinguish it from the strictly DNA-based group I intron homing and strictly RNA-based group II intron transposition.
- 20Saldanha, R.; Chen, B.; Wank, H.; Matsuura, M.; Edwards, J.; Lambowitz, A. M. RNA and protein catalysis in group II intron splicing and mobility reactions using purified components. Biochemistry 1999, 38, 9069– 9083, DOI: 10.1021/bi982799lGoogle Scholar20https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXjvVCqtro%253D&md5=2136004d61355a78d3ad72c72f646545RNA and Protein Catalysis in Group II Intron Splicing and Mobility Reactions Using Purified ComponentsSaldanha, Roland; Chen, Bing; Wank, Herbert; Matsuura, Manabu; Edwards, Judy; Lambowitz, Alan M.Biochemistry (1999), 38 (28), 9069-9083CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)Group II introns encode proteins with reverse transcriptase activity. These proteins also promote RNA splicing (maturase activity) and then, with the excised intron, form a site-specific DNA endonuclease that promotes intron mobility by reverse splicing into DNA followed by target DNA-primed reverse transcription. Here, we used an Escherichia coli expression system for the Lactococcus lactis group II intron Ll.LtrB to show that the intron-encoded protein (LtrA) alone is sufficient for maturase activity, and that RNP particles contg. only the LtrA protein and excised intron RNA have site-specific DNA endonuclease and target DNA-primed reverse transcriptase activity. Detailed anal. of the splicing reaction indicates that LtrA is an intron-specific splicing factor that binds to unspliced precursor RNA with a Kd of ≤0.12 pM at 30°. This binding occurs in a rapid bimol. reaction, which is followed by a slower step, presumably an RNA conformational change, required for splicing to occur. Our results constitute the first biochem. anal. of protein-dependent splicing of a group II intron and demonstrate that a single intron-encoded protein can interact with the intron RNA to carry out a coordinated series of reactions leading to splicing and mobility.
- 21Matsuura, M.; Noah, J. W.; Lambowitz, A. M. Mechanism of maturase-promoted group II intron splicing. EMBO J. 2001, 20, 7259– 7270, DOI: 10.1093/emboj/20.24.7259Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XpsF2h&md5=18a1b077ef7c4155b12d2fc0bef65540Mechanism of maturase-promoted group II intron splicingMatsuura, Manabu; Noah, James W.; Lambowitz, Alan M.EMBO Journal (2001), 20 (24), 7259-7270CODEN: EMJODG; ISSN:0261-4189. (Oxford University Press)Mobile group II introns encode reverse transcriptases that also function as intron-specific splicing factors (maturases). We showed previously that the reverse transcriptase/maturase encoded by the Lactococcus lactis Ll.LtrB intron has a high affinity binding site at the beginning of its own coding region in an idiosyncratic structure, DIVa. Here, we identify potential secondary binding sites in conserved regions of the catalytic core and show via chem. modification expts. that binding of the maturase induces the formation of key tertiary interactions required for RNA splicing. The interaction with conserved as well as idiosyneratic regions explains how maturases in some organisms could evolve into general group II intron splicing factors, potentially mirroring a key step in the evolution of spliceosomal introns.
- 22Lambowitz, A. M.; Zimmerly, S. Group II introns: Mobile ribozymes that invade DNA. Cold Spring Harbor Perspect. Biol. 2011, 3, a003616 DOI: 10.1101/cshperspect.a003616Google ScholarThere is no corresponding record for this reference.
- 23Perutka, J.; Wang, W.; Goerlitz, D.; Lambowitz, A. M. Use of Computer-designed Group II Introns to Disrupt Escherichia coli DExH/D-box Protein and DNA Helicase Genes. J. Mol. Biol. 2004, 336, 421– 439, DOI: 10.1016/j.jmb.2003.12.009Google Scholar23https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2cXntVKrsw%253D%253D&md5=d7cde4aede8256f94d4df7eeed98ed2bUse of Computer-designed Group II Introns to Disrupt Escherichia coli DExH/D-box Protein and DNA Helicase GenesPerutka, Jiri; Wang, Wenjun; Goerlitz, David; Lambowitz, Alan M.Journal of Molecular Biology (2004), 336 (2), 421-439CODEN: JMOBAK; ISSN:0022-2836. (Elsevier)Mobile group II introns are site-specific retroelements that use a novel mobility mechanism in which the excised intron RNA inserts directly into a DNA target site and is then reverse transcribed by the assocd. intron-encoded protein. Because the DNA target site is recognized primarily by base-pairing of the intron RNA with only a small no. of positions recognized by the protein, it has been possible to develop group II introns into a new type of gene targeting vector ("targetron"), which can be reprogrammed to insert into desired DNA targets simply by modifying the intron RNA. Here, we used databases of retargeted Lactococcus lactis Ll.LtrB group II introns and a compilation of nucleotide frequencies at active target sites to develop an algorithm that predicts optimal Ll.LtrB intron-insertion sites and designs primers for modifying the intron to insert into those sites. In a test of the algorithm, we designed one or two targetrons to disrupt each of 28 Escherichia coli genes encoding DExH/D-box and DNA helicase-related proteins and tested for the desired disruptants by PCR screening of 100 colonies. In 21 cases, we obtained disruptions at frequencies of 1-80% without selection, and in six other cases, where disruptants were not identified in the initial PCR screen, we readily obtained specific disruptions by using the same targetrons with a retrotransposition-activated selectable marker. Only one DExH/D-box protein gene, secA, which was known to be essential, did not give viable disruptants. The apparent dispensability of DExH/D-box proteins in E. coli contrasts with the situation in yeast, where the majority of such proteins are essential. The methods developed here should permit the rapid and efficient disruption of any bacterial gene, the computational anal. provides new insight into group II intron target site recognition, and the set of E. coli DExH/D-box protein and DNA helicase disruptants should be useful for analyzing the function of these proteins.
- 24Heap, J. T.; Pennington, O. J.; Cartman, S. T.; Carter, G. P.; Minton, N. P. The ClosTron: A universal gene knock-out system for the genus Clostridium. J. Microbiol. Methods 2007, 70, 452– 464, DOI: 10.1016/j.mimet.2007.05.021Google Scholar24https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXpsFKjsLk%253D&md5=5ad07920edc8de3b8f622522223016aaThe ClosTron: A universal gene knock-out system for the genus ClostridiumHeap, John T.; Pennington, Oliver J.; Cartman, Stephen T.; Carter, Glen P.; Minton, Nigel P.Journal of Microbiological Methods (2007), 70 (3), 452-464CODEN: JMIMDQ; ISSN:0167-7012. (Elsevier B.V.)Progress in exploiting clostridial genome information has been severely impeded by a general lack of effective methods for the directed inactivation of specific genes. Those few mutants that have been generated have been almost exclusively derived by single crossover integration of a replication-deficient or defective plasmid by homologous recombination. The mutants created are therefore unstable. Here we have adapted a mutagenesis system based on the mobile group II intron from the ltrB gene of Lactococcus lactis (Ll.ltrB) to function in clostridial hosts. Integrants are readily selected on the basis of acquisition of resistance to erythromycin, and are generated from start to finish in as little as 10 to 14 days. Unlike single crossover plasmid integrants, the mutants are extremely stable. The system has been used to make 6 mutants of Clostridium acetobutylicum and 5 of Clostridium difficile, exceeding the no. of published mutants ever generated in these species. Genes have also been inactivated for the first time in Clostridium botulinum and Clostridium sporogenes, suggesting the system will be universally applicable to the genus. The procedure is highly efficient and reproducible, and should revolutionize functional genomic studies in clostridia.
- 25Akhtar, P.; Khan, S. A. Two independent replicons can support replication of the anthrax toxin-encoding plasmid pXO1 of Bacillus anthracis. Plasmid 2012, 67, 111– 117, DOI: 10.1016/j.plasmid.2011.12.012Google Scholar25https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC38Xks12jtrY%253D&md5=e29997ac16445fd79197fc6dbd55ccb6Two independent replicons can support replication of the anthrax toxin-encoding plasmid pXO1 of Bacillus anthracisAkhtar, Parvez; Khan, Saleem A.Plasmid (2012), 67 (2), 111-117CODEN: PLSMDX; ISSN:0147-619X. (Elsevier)The large pXO1 plasmid (181.6 kb) of Bacillus anthracis encodes the anthrax toxin proteins. Previous studies have shown that two sep. regions of pXO1 can support replication of pXO1 miniplasmids when introduced into plasmid-less strains of this organism. No information is currently available on the ability of the above two replicons, termed RepX and ORFs 14/16 replicons, to support replication of the full-length pXO1 plasmid. We generated mutants of the full-length pXO1 plasmid in which either the RepX or the ORFs 14/16 replicon was inactivated by TargeTron insertional mutagenesis. Plasmid pXO1 derivs. contg. only the RepX or the ORFs 14/16 replicon were able to replicate when introduced into a plasmid-less B. anthracis strain. Plasmid copy no. anal. showed that the ORFs 14/16 replicon is more efficient than the RepX replicon. Our studies demonstrate that both the RepX and ORFs 14/16 replicons can independently support the replication of the full-length pXO1 plasmid.
- 26Frazier, C. L.; San Filippo, J.; Lambowitz, A. M.; Mills, D. A. Genetic manipulation of Lactococcus lactis by using targeted group II introns: Generation of stable insertions without selection. Appl. Environ. Microbiol. 2003, 69, 1121– 1128, DOI: 10.1128/AEM.69.2.1121-1128.2003Google Scholar26https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXhtF2it7w%253D&md5=70f040e4515aded320ec667e68c53cceGenetic manipulation of Lactococcus lactis by using targeted group II introns: Generation of stable insertions without selectionFrazier, Courtney L.; San Filippo, Joseph; Lambowitz, Alan M.; Mills, David A.Applied and Environmental Microbiology (2003), 69 (2), 1121-1128CODEN: AEMIDF; ISSN:0099-2240. (American Society for Microbiology)Despite their com. importance, there are relatively few facile methods for genomic manipulation of the lactic acid bacteria. Here, the lactococcal group II intron, Ll.ltrB, was targeted to insert efficiently into genes encoding malate decarboxylase (mleS) and tetracycline resistance (tetM) within the Lactococcus lactis genome. Integrants were readily identified and maintained in the absence of a selectable marker. Since splicing of the Ll.ltrB intron depends on the intron-encoded protein, targeted invasion with an intron lacking the intron open reading frame disrupted TetM and MleS function, and MleS activity could be partially restored by expressing the intron-encoded protein in trans. Restoration of splicing from intron variants lacking the intron-encoded protein illustrates how targeted group II introns could be used for conditional expression of any gene. Furthermore, the modified Ll.ltrB intron was used to sep. deliver a phage resistance gene (abiD) and a tetracycline resistance marker (tetM) into mleS, without the need for selection to drive the integration or to maintain the integrant. Our findings demonstrate the utility of targeted group II introns as a potential food-grade mechanism for delivery of industrially important traits into the genomes of lactococci.
- 27Plante, I.; Cousineau, B. Restriction for gene insertion within the Lactococcus lactis Ll.LtrB group II intron. RNA 2006, 12, 1980– 1992, DOI: 10.1261/rna.193306Google Scholar27https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XhtFygsL%252FF&md5=12308f25b4b51b486da6ce6398af0b73Restriction for gene insertion within the Lactococcus lactis Ll.LtrB group II intronPlante, Isabelle; Cousineau, BenoitRNA (2006), 12 (11), 1980-1992CODEN: RNARFU; ISSN:1355-8382. (Cold Spring Harbor Laboratory Press)The Ll.LtrB intron, from the low G+C gram-pos. bacterium Lactococcus lactis, was the first bacterial group II intron shown to splice and mobilize in vivo. The detailed retrohoming and retrotransposition pathways of Ll.LtrB were studied in both L. lactis and Escherichia coli. This bacterial retroelement has many features that would make it a good gene delivery vector. Here we report that the mobility efficiency of Ll.LtrB expressing LtrA in trans is only slightly affected by the insertion of fragments <100 nucleotides within the loop region of domain IV. In contrast, Ll.LtrB mobility efficiency is drastically decreased by the insertion of foreign sequences >1 kb. We demonstrate that the inhibitory effect caused by the addn. of expression cassettes on Ll.LtrB mobility efficiency is not sequence specific, and not due to the expression, or the toxicity, of the cargo genes. Using genetic screens, we demonstrate that in order to maintain intron mobility, the loop region of domain IV, more specifically domain IVb, is by far the best region to insert foreign sequences within Ll.LtrB. Poisoned primer extension and Northern blot analyses reveal that Ll.LtrB constructs harboring cargo sequences splice less efficiently, and show a significant redn. in lariat accumulation in L. lactis. This suggests that cargo-contg. Ll.LtrB variants are less stable. These results reveal the potential, yet limitations, of the Ll.LtrB group II intron to be used as a gene delivery vector, and validate the random insertion approach described in this study to create cargo-contg. Ll.LtrB variants that are mobile.
- 28Rawsthorne, H.; Turner, K. N.; Mills, D. A. Multicopy integration of heterologous genes, using the lactococcal group II intron targeted to bacterial insertion sequences. Appl. Environ. Microbiol. 2006, 72, 6088– 6093, DOI: 10.1128/AEM.02992-05Google Scholar28https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XpvVKjt70%253D&md5=c1d1e67be17b38c6d7a699b3b11ca654Multicopy integration of heterologous genes, using the lactococcal group II intron targeted to bacterial insertion sequencesRawsthorne, Helen; Turner, Kevin N.; Mills, David A.Applied and Environmental Microbiology (2006), 72 (9), 6088-6093CODEN: AEMIDF; ISSN:0099-2240. (American Society for Microbiology)Group II introns are mobile genetic elements that can be redirected to invade specific genes. Here the authors describe the use of the lactococcal group II intron, Ll.ltrB, to achieve multicopy delivery of heterologous genes into the genome of Lactococcus lactis IL1403-UCD without the need for selectable markers. Ll.ltrB was retargeted to invade three transposase genes, the tra gene found in IS904 (tra904), tra981, and tra983, of which 9, 10, and 14 copies, resp., were present in IL1403-UCD. Intron invasion of tra904, tra981, and tra983 allele groups occurred at high frequencies, and individual segregants possessed anywhere from one to nine copies of intron in the resp. tra alleles. To achieve multicopy delivery of a heterologous gene, a green fluorescent protein (GFP) marker was cloned into the tra904-targeted Ll.ltrB, and the resultant intron (Ll.ltrB::GFP) was induced to invade the L. lactis tra904 alleles. Segregants possessing Ll.ltrB::GFP in three, four, five, six, seven, and eight copies in different tra904 alleles were obtained. In general, increasing the chromosomal copy no. of Ll.ltrB::GFP resulted in strains expressing successively higher levels of GFP. However, strains possessing the same no. of Ll.ltrB::GFP copies within different sets of tra904 alleles exhibited differential GFP expression, and segregants possessing seven or eight copies of Ll.ltrB::GFP grew poorly upon induction, suggesting that GFP expression from certain combinations of alleles was detrimental. The highest level of GFP expression was obsd. from a specific six-copy variant that produced GFP at a level analogous to that obtained with a multicopy plasmid. In addn., the high level of GFP expression was stable for over 120 generations. This work demonstrates that stable multicopy integration of heterologous genes can be readily achieved in bacterial genomes with group II intron delivery by targeting repeated elements.
- 29Yao, J.; Zhong, J.; Fang, Y.; Geisinger, E.; Novick, R. P.; Lambowitz, A. M. Use of targetrons to disrupt essential and nonessential genes in Staphylococcus aureus reveals temperature sensitivity of Ll.LtrB group II intron splicing. RNA 2006, 12, 1271– 1281, DOI: 10.1261/rna.68706Google Scholar29https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28Xms1Cqu78%253D&md5=4e089632c8b7390ccce87c3513cbe723Use of targetrons to disrupt essential and nonessential genes in Staphylococcus aureus reveals temperature sensitivity of Ll.LtrB group II intron splicingYao, Jun; Zhong, Jin; Fang, Yuan; Geisinger, Edward; Novick, Richard P.; Lambowitz, Alan M.RNA (2006), 12 (7), 1271-1281CODEN: RNARFU; ISSN:1355-8382. (Cold Spring Harbor Laboratory Press)We show that a targetron based on the Lactococcus lactis Ll.LtrB group II intron can be used for efficient chromosomal gene disruption in the human pathogen Staphylococcus aureus. Targetrons expressed from derivs. of vector pCN37, which uses a cadmium-inducible promoter, or pCN39, a deriv. of pCN37 with a temp.-sensitive replicon, gave site-specific disruptants of the hsa and seb genes in 37%-100% of plated colonies without selection. To disrupt hsa, an essential gene, we used a group II intron that integrates in the sense orientation relative to target gene transcription and thus could be removed by RNA splicing, enabling the prodn. of functional HSa protein. We show that because splicing of the Ll.LtrB intron by the intron-encoded protein is temp.-sensitive, this method yields a conditional hsa disruptant that grows at 32° but not 43°. The temp. sensitivity of the splicing reaction suggests a general means of obtaining one-step conditional disruptions in any organism. In nature, temp. sensitivity of group II intron splicing could limit the temp. range of an organism contg. a group II intron inserted in an essential gene.
- 30Karberg, M.; Guo, H.; Zhong, J.; Coon, R.; Perutka, J.; Lambowitz, A. M. Group II introns as controllable gene targeting vectors for genetic manipulation of bacteria. Nat. Biotechnol. 2001, 19, 1162– 1167, DOI: 10.1038/nbt1201-1162Google Scholar30https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXptV2qsrw%253D&md5=3dad957336fc3ad3a37f4316d019bf4aGroup II introns as controllable gene targeting vectors for genetic manipulation of bacteriaKarberg, Michael; Guo, Huatao; Zhong, Jin; Coon, Robert; Perutka, Jiri; Lambowitz, Alan M.Nature Biotechnology (2001), 19 (12), 1162-1167CODEN: NABIF9; ISSN:1087-0156. (Nature America Inc.)Mobile group II introns can be retargeted to insert into virtually any desired DNA target. Here the authors show that retargeted group II introns can be used for highly specific chromosomal gene disruption in Escherichia coli and other bacteria at frequencies of 0.1-22%. Furthermore, the introns can be used to introduce targeted chromosomal breaks, which can be repaired by transformation with a homologous DNA fragment, enabling the introduction of point mutations. Because of their wide host range, mobile group II introns should be useful for genetic engineering and functional genomics in a wide variety of bacteria.
- 31García-Rodríguez, F. M.; Hernańdez-Gutiérrez, T.; Diáz-Prado, V.; Toro, N. Use of the computer-retargeted group II intron RmInt1 of Sinorhizobium meliloti for gene targeting. RNA Biol. 2014, 11, 391– 401, DOI: 10.4161/rna.28373Google Scholar31https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC2crlsFehsA%253D%253D&md5=49269dde7cb61dfd2adf91c2aa0d3971Use of the computer-retargeted group II intron RmInt1 of Sinorhizobium meliloti for gene targetingGarcia-Rodriguez Fernando M; Hernandez-Gutierrez Teresa; Diaz-Prado Vanessa; Toro NicolasRNA biology (2014), 11 (4), 391-401 ISSN:.Gene-targeting vectors derived from mobile group II introns capable of forming a ribonucleoprotein (RNP) complex containing excised intron lariat RNA and an intron-encoded protein (IEP) with reverse transcriptase (RT), maturase, and endonuclease (En) activities have been described. RmInt1 is an efficient mobile group II intron with an IEP lacking the En domain. We performed a comprehensive study of the rules governing RmInt1 target site recognition based on selection experiments with donor and recipient plasmid libraries, with randomization of the elements of the intron RNA involved in target recognition and the wild-type target site. The data obtained were used to develop a computer algorithm for identifying potential RmInt1 targets in any DNA sequence. Using this algorithm, we modified RmInt1 for the efficient recognition of DNA target sites at different locations in the Sinorhizobium meliloti chromosome. The retargeted RmInt1 integrated efficiently into the chromosome, regardless of the location of the target gene. Our results suggest that RmInt1 could be efficiently adapted for gene targeting.
- 32Cousineau, B.; Smith, D.; Lawrence-Cavanagh, S.; Mueller, J. E.; Yang, J.; Mills, D.; Manias, D.; Dunny, G.; Lambowitz, A. M.; Belfort, M. Retrohoming of a bacterial group II intron: Mobility via complete reverse splicing, independent of homologous DNA recombination. Cell 1998, 94, 451– 462, DOI: 10.1016/S0092-8674(00)81586-XGoogle Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXlslOktb8%253D&md5=ebbed8461cc1dadf9c870f01354bac6dRetrohoming of a bacterial group II intron: mobility via complete reverse splicing, independent of homologous DNA recombinationCousineau, Benoit; Smith, Dorie; Lawrence-Cavanagh, Stacey; Mueller, John E.; Yang, Jian; Mills, David; Dunny, Gary; Lambowitz, Alan M.; Belfort, MarleneCell (Cambridge, Massachusetts) (1998), 94 (4), 451-462CODEN: CELLB5; ISSN:0092-8674. (Cell Press)The mobile group II intron of Lactococcus lactis, LI.LtrB, provides the opportunity to analyze the homing pathway in genetically tractable bacterial systems. Here, we show that LI.LtrB mobility occurs by an RNA-based retrohoming mechanism in both Escherichia coli and L. lactis. Surprisingly, retrohoming occurs efficiently in the absence of RecA function, with a relaxed requirement for flanking exon homol. and without coconversion of exon markers. These results lead to a model for bacterial retrohoming in which the intron integrates into recipient DNA by complete reverse splicing and serves as the template for cDNA synthesis. The retrohoming reaction is completed in unprecedented fashion by a DNA repair event that is independent of homologous recombination between the alleles. Thus, LI.LtrB has many features of retrotransposons, with practical and evolutionary implications.
- 33Velázquez, E.; Lorenzo, V. D.; Al-Ramahi, Y. Recombination-Independent Genome Editing through CRISPR/Cas9-Enhanced TargeTron Delivery. ACS Synth. Biol. 2019, 8, 2186– 2193, DOI: 10.1021/acssynbio.9b00293Google Scholar33https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhsF2ntbzN&md5=f06edd98e918993b500d50caa1978a9cRecombination-Independent Genome Editing through CRISPR/Cas9-Enhanced TargeTron DeliveryVelazquez, Elena; Lorenzo, Victor de; Al-Ramahi, YamalACS Synthetic Biology (2019), 8 (9), 2186-2193CODEN: ASBCD6; ISSN:2161-5063. (American Chemical Society)Group II introns were developed time ago as tools for the construction of knockout mutants in a wide range of organisms, ranging from Gram-pos. and Gram-neg. bacteria to human cells. Utilizing these introns is advantageous because they are independent of the host's DNA recombination machinery, they can carry heterologous sequences (and thus be used as vehicles for gene delivery), and they can be easily retargeted for subsequent insertions of addnl. genes at the user's will. Alas, the use of this platform has been limited, as insertion efficiencies greatly change depending on the target sites and cannot be predicted a priori. Moreover, the ability of introns to perform their own splicing and integration is compromised when they carry foreign sequences. To overcome these limitations, we merged the group II intron-based TargeTron system with CRISPR/Cas9 counter-selection. To this end, we first engineered a new group-II intron by replacing the retrotransposition-activated selectable marker (RAM) with ura3 and retargeting it to a new site in the lacZ gene of E. coli. Then, we showed proved that directing CRISPR/Cas9 towards the wild-type sequences dramatically increased the chances of finding clones that integrated the retrointron into the target lacZ sequence. The CRISPR-Cas9 counter selection strategy presented herein thus overcomes a major limitation that has prevented the use of group II introns as devices for gene delivery and genome editing at large in a recombination-independent fashion.
- 34Barrangou, R.; Fremaux, C.; Deveau, H.; Richards, M.; Boyaval, P.; Moineau, S.; Romero, D. A.; Horvath, P. CRISPR provides acquired resistance against viruses in prokaryotes. Science 2007, 315, 1709– 1712, DOI: 10.1126/science.1138140Google Scholar34https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXjtlWntb8%253D&md5=b88c0100d2c0469213afda20e47c39cdCRISPR Provides Acquired Resistance Against Viruses in ProkaryotesBarrangou, Rodolphe; Fremaux, Christophe; Deveau, Helene; Richards, Melissa; Boyaval, Patrick; Moineau, Sylvain; Romero, Dennis A.; Horvath, PhilippeScience (Washington, DC, United States) (2007), 315 (5819), 1709-1712CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)Clustered regularly interspaced short palindromic repeats (CRISPR) are a distinctive feature of the genomes of most Bacteria and Archaea and are thought to be involved in resistance to bacteriophages. We found that, after viral challenge, bacteria integrated new spacers derived from phage genomic sequences. Removal or addn. of particular spacers modified the phage-resistance phenotype of the cell. Thus, CRISPR, together with assocd. cas genes, provided resistance against phages, and resistance specificity is detd. by spacer-phage sequence similarity.
- 35Jiang, W.; Bikard, D.; Cox, D.; Zhang, F.; Marraffini, L. A. RNA-guided editing of bacterial genomes using CRISPR-Cas systems. Nat. Biotechnol. 2013, 31, 233– 239, DOI: 10.1038/nbt.2508Google Scholar35https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXhsFCkurY%253D&md5=ea3bbfda133b51e71b114de2c9ceabfdRNA-guided editing of bacterial genomes using CRISPR-Cas systemsJiang, Wenyan; Bikard, David; Cox, David; Zhang, Feng; Marraffini, Luciano A.Nature Biotechnology (2013), 31 (3), 233-239CODEN: NABIF9; ISSN:1087-0156. (Nature Publishing Group)Here we use the clustered, regularly interspaced, short palindromic repeats (CRISPR)-assocd. Cas9 endonuclease complexed with dual-RNAs to introduce precise mutations in the genomes of Streptococcus pneumoniae and Escherichia coli. The approach relies on dual-RNA:Cas9-directed cleavage at the targeted genomic site to kill unmutated cells and circumvents the need for selectable markers or counter-selection systems. We reprogram dual-RNA:Cas9 specificity by changing the sequence of short CRISPR RNA (crRNA) to make single- and multinucleotide changes carried on editing templates. Simultaneous use of two crRNAs enables multiplex mutagenesis. In S. pneumoniae, nearly 100% of cells that were recovered using our approach contained the desired mutation, and in E. coli, 65% that were recovered contained the mutation, when the approach was used in combination with recombineering. We exhaustively analyze dual-RNA:Cas9 target requirements to define the range of targetable sequences and show strategies for editing sites that do not meet these requirements, suggesting the versatility of this technique for bacterial genome engineering.
- 36Aparicio, T.; de Lorenzo, V.; Martínez-García, E. CRISPR/Cas9-Based Counterselection Boosts Recombineering Efficiency in Pseudomonas putida. Biotechnol. J. 2018, 13, 1700161 DOI: 10.1002/biot.201700161Google ScholarThere is no corresponding record for this reference.
- 37Benedetti, I.; Nikel, P. I.; de Lorenzo, V. Data on the standardization of a cyclohexanone-responsive expression system for Gram-negative bacteria. Data Brief 2016, 6, 738– 744, DOI: 10.1016/j.dib.2016.01.022Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC28jgt1Onug%253D%253D&md5=5c2853ffbcc263263df26ad415e9484aData on the standardization of a cyclohexanone-responsive expression system for Gram-negative bacteriaBenedetti Ilaria; Nikel Pablo I; de Lorenzo VictorData in brief (2016), 6 (), 738-44 ISSN:2352-3409.Engineering of robust microbial cell factories requires the use of dedicated genetic tools somewhat different from those traditionally used for laboratory-adapted microorganisms. We have edited and formatted the ChnR/P chnB regulatory node of Acinetobacter johnsonii to ease the targeted engineering of ectopic gene expression in Gram-negative bacteria. The proposed compositional standard was thoroughly verified with a monomeric and superfolder green fluorescent protein (msf•GFP) in Escherichia coli. The expression data presented reflect a tightly controlled transcription initiation signal in response to cyclohexanone. Data in this article are related to the research paper "Genetic programming of catalytic Pseudomonas putida biofilms for boosting biodegradation of haloalkanes" [1].
- 38Yao, J.; Lambowitz, A. M. Gene targeting in Gram-negative bacteria by use of a mobile group II intron (“targetron”) expressed from a broad-host-range vector. Appl. Environ. Microbiol. 2007, 73, 2735– 2743, DOI: 10.1128/AEM.02829-06Google Scholar38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXkvFeqsbc%253D&md5=136a1d18e1873775be03c8940174d9feGene targeting in gram-negative bacteria by use of a mobile group II intron ("targetron") expressed from a broad-host-range vectorYao, Jun; Lambowitz, Alan M.Applied and Environmental Microbiology (2007), 73 (8), 2735-2743CODEN: AEMIDF; ISSN:0099-2240. (American Society for Microbiology)Mobile group II introns ("targetrons") can be programmed for insertion into virtually any desired DNA target with high frequency and specificity. Here, we show that targetrons expressed via an m-toluic acid-inducible promoter from a broad-host-range vector contg. an RK2 minireplicon can be used for efficient gene targeting in a variety of gram-neg. bacteria, including Escherichia coli, Pseudomonas aeruginosa, and Agrobacterium tumefaciens. Targetrons expressed from donor plasmids introduced by electroporation or conjugation yielded targeted disruptions at frequencies of 1 to 58% of screened colonies in the E. coli lacZ, P. aeruginosa pqsA and pqsH, and A. tumefaciens aopB and chvI genes. The development of this broad-host-range system for targetron expression should facilitate gene targeting in many bacteria.
- 39Martínez-García, E.; Aparicio, T.; Goñi-Moreno, A.; Fraile, S.; De Lorenzo, V. SEVA 2.0: An update of the Standard European Vector Architecture for de-/re-construction of bacterial functionalities. Nucleic Acids Res. 2015, 43, D1183– D1189, DOI: 10.1093/nar/gku1114Google Scholar39https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhtVymtLfL&md5=8152ca1366af4c0fe1d99435b137bbcbSEVA 2.0: an update of the standard European vector architecture for de-/re-construction of bacterial functionalitiesMartinez-Garcia, Esteban; Aparicio, Tomas; Goni-Moreno, Angel; Fraile, Sofia; de Lorenzo, VictorNucleic Acids Research (2015), 43 (D1), D1183-D1189CODEN: NARHAD; ISSN:0305-1048. (Oxford University Press)A review. The Std. European Vector Architecture 2.0 database (SEVA-DB 2.0) is an improved and expanded version of the platform released in 2013 aimed at assisting the choice of optimal genetic tools for deconstructing and re-constructing complex prokaryotic phenotypes. By adopting simple compositional rules, the SEVA std. facilitates combinations of functional DNA segments that ease both the anal. and the engineering of diverse Gram-neg. bacteria for fundamental or biotechnol. purposes. The large no. of users of the SEVA-DB during its first two years of existence has resulted in a valuable feedback that we have exploited for fixing DNA sequence errors, improving the nomenclature of the SEVA plasmids, expanding the vector collection, adding new features to the web interface and encouraging contributions of materials from the community of users. The SEVA platform is also adopting the Synthetic Biol. Open Language (SBOL) for electronic-like description of the constructs available in the collection and their interfacing with genetic devices developed by other Synthetic Biol. communities. We advocate the SEVA format as one interim asset for the ongoing transition of genetic design of microorganisms from being a trial-and-error endeavor to become an authentic engineering discipline.
- 40Martínez-García, E.; Goñi-Moreno, A.; Bartley, B.; McLaughlin, J.; Sánchez-Sampedro, L.; Pascual Del Pozo, H.; Prieto Hernández, C.; Marletta, A. S.; De Lucrezia, D.; Sánchez-Fernández, G.; Fraile, S.; De Lorenzo, V. SEVA 3.0: An update of the Standard European Vector Architecture for enabling portability of genetic constructs among diverse bacterial hosts. Nucleic Acids Res. 2020, 48, D1164– D1170, DOI: 10.1093/nar/gkz1024Google Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXhs1GltrbP&md5=4f5a11f9bf0a49d09a3ff86bcfdd1252SEVA 3.0: an update of the standard European vector architecture for enabling portability of genetic constructs among diverse bacterial hostsMartinez-Garcia, Esteban; Goni-Moreno, Angel; Bartley, Bryan; McLaughlin, James; Sanchez-Sampedro, Lucas; del Pozo, Hector Pascual; Hernandez, Clara Prieto; Marletta, Ada Serena; De Lucrezia, Davide; Sanchez-Fernandez, Guzman; Fraile, Sofia; de Lorenzo, VictorNucleic Acids Research (2020), 48 (D1), D1164-D1170CODEN: NARHAD; ISSN:1362-4962. (Oxford University Press)A review. The Std. European Vector Architecture 3.0 database is the update of the platform launched in 2013 both as a web-based resource and as a material repository of formatted genetic tools (mostly plasmids) for anal., construction and deployment of complex bacterial phenotypes. The period between the first version of SEVA-DB and the present time has witnessed several tech., computational and conceptual advances in genetic/genomic engineering of prokaryotes that have enabled upgrading of the utilities of the updated database. Novelties include not only a more user-friendly web interface and many more plasmid vectors, but also new links of the plasmids to advanced bioinformatic tools. These provide an intuitive visualization of the constructs at stake and a range of virtual manipulations of DNA segments that were not possible before. Finally, the list of canonical SEVA plasmids is available in machine-readable SBOL (Synthetic Biol. Open Language) format. This ensures interoperability with other platforms and affords simulations of their behavior under different in vivo conditions. We argue that the SEVA-DB will remain a useful resource for extending Synthetic Biol. approaches towards non-std. bacterial species as well as genetically programming new prokaryotic chassis for a suite of fundamental and biotechnol. endeavours.
- 41Zhong, J.; Karberg, M.; Lambowitz, A. M. Targeted and random bacterial gene disruption using a group II intron (targetron) vector containing a retrotransposition-activated selectable marker. Nucleic Acids Res. 2003, 31, 1656– 1664, DOI: 10.1093/nar/gkg248Google Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3sXit1yks7g%253D&md5=9ddea609f3357b85ce578dbb1967c657Targeted and random bacterial gene disruption using a group II intron (targetron) vector containing a retrotransposition-activated selectable markerZhong, Jin; Karberg, Michael; Lambowitz, Alan M.Nucleic Acids Research (2003), 31 (6), 1656-1664CODEN: NARHAD; ISSN:0305-1048. (Oxford University Press)Mobile group II introns have been used to develop a novel class of gene targeting vectors, targetrons, which employ base pairing for DNA target recognition and can thus be programmed to insert into any desired target DNA. Here, we have developed a targetron contg. a retrotransposition-activated selectable marker (RAM), which enables one-step bacterial gene disruption at near 100% efficiency after selection. The targetron can be generated via PCR without cloning, and after intron integration, the marker gene can be excised by recombination between flanking Flp recombinase sites, enabling multiple sequential disruptions. We also show that a RAM-targetron with randomized target site recognition sequences yields single insertions throughout the Escherichia coli genome, creating a gene knockout library. Anal. of the randomly selected insertion sites provides further insight into group II intron target site recognition rules. It also suggests that a subset of retrohoming events may occur by using a primer generated during DNA replication, and reveals a previously unsuspected bias for group II intron insertion near the chromosome replication origin. This insertional bias likely reflects at least in part the higher copy no. of origin proximal genes, but interaction with the replication machinery or other features of DNA structure or packaging may also contribute.
- 42Studier, F. W.; Moffatt, B. A. Use of bacteriophage T7 RNA polymerase to direct selective high-level expression of cloned genes. J. Mol. Biol. 1986, 189, 113– 130, DOI: 10.1016/0022-2836(86)90385-2Google Scholar42https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL28XktlKrsr4%253D&md5=3219300bc2f640fe9830c7518eb99bfcUse of bacteriophage T7 RNA polymerase to direct selective high-level expression of cloned genesStudier, F. William; Moffatt, Barbara A.Journal of Molecular Biology (1986), 189 (1), 113-30CODEN: JMOBAK; ISSN:0022-2836.A gene expression system based on phage T7 RNA polymerase [9014-24-8] was developed. T7 RNA polymerase is highly selective for its own promoters, which do not occur naturally in Escherichia coli. A relatively small amt. of T7 RNA polymerase provided from a cloned copy of T7 gene 1 is sufficient to direct high-level transcription from a T7 promoter in a multicopy plasmid. Such transcription can proceed several times around the plasmid without terminating and can be so active that transcription by E. coli RNA polymerase is greatly decreased. When a cleavage site for RNase III is introduced, discrete RNAs of plasmid length can accumulate. The natural transcription terminator from T7 DNA also works effectively in the plasmid. Both the rate of synthesis and the accumulation of RNA directed by T7 RNA polymerase can reach levels comparable with those for rRNAs in a normal cell. These high levels of accumulation suggest that the RNAs are relatively stable, perhaps in part because their great length and(or) stem-and-loop structures at their 3' ends help to protect them against exonucleolytic degrdn. Apparently, a specific mRNA produced by T7 RNA polymerase can rapidly sat. the translational machinery of E. coli, so that the rate of protein synthesis from such an mRNA will depend primarily on the efficiency of its translation. When the mRNA is efficiently translated, a target protein can accumulate to >50% of the total cell protein in ≤3 h. Two ways were used to deliver active T7 RNA polymerase to the cell: (1) infection by a λ deriv. that carried gene 1; or (2) induction of a chromosomal copy of gene 1 under control of the lacUV5 promoter. When gene 1 is delivered by infection, very toxic target genes can be maintained silently in the cell until T7 RNA polymerase is introduced, when they rapidly become expressed at high levels. When gene 1 is resident in the chromosome, even the very low basal levels of T7 RNA polymerase present in the uninduced cell can prevent the establishment of plasmids carrying toxic target genes, or make the plasmid unstable. But if the target plasmid can be maintained, induction of chromosomal gene 1 can be a convenient way to produce large amts. of target RNA and(or) protein. T7 RNA polymerase seems to be capable of transcribing almost any DNA linked to a T7 promoter, so the T7 expression system should be capable of transcribing almost any gene or its complement in E. coli. Comparable T7 expression systems can be developed in other types of cells.
- 43Dubendorf, J. W.; Studier, F. W. Controlling basal expression in an inducible T7 expression system by blocking the target T7 promoter with lac repressor. J. Mol. Biol. 1991, 219, 45– 59, DOI: 10.1016/0022-2836(91)90856-2Google ScholarThere is no corresponding record for this reference.
- 44Angius, F.; Ilioaia, O.; Amrani, A.; Suisse, A.; Rosset, L.; Legrand, A.; Abou-Hamdan, A.; Uzan, M.; Zito, F.; Miroux, B. A novel regulation mechanism of the T7 RNA polymerase based expression system improves overproduction and folding of membrane proteins. Sci. Rep. 2018, 8, 8572 DOI: 10.1038/s41598-018-26668-yGoogle Scholar44https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC1Mbkt12isg%253D%253D&md5=0256960f3095dbbacbc5be9d233b3c38A novel regulation mechanism of the T7 RNA polymerase based expression system improves overproduction and folding of membrane proteinsAngius Federica; Ilioaia Oana; Amrani Amira; Suisse Annabelle; Rosset Lindsay; Legrand Amelie; Abou-Hamdan Abbas; Uzan Marc; Zito Francesca; Miroux Bruno; Suisse Annabelle; Abou-Hamdan AbbasScientific reports (2018), 8 (1), 8572 ISSN:.Membrane protein (MP) overproduction is one of the major bottlenecks in structural genomics and biotechnology. Despite the emergence of eukaryotic expression systems, bacteria remain a cost effective and powerful tool for protein production. The T7 RNA polymerase (T7RNAP)-based expression system is a successful and efficient expression system, which achieves high-level production of proteins. However some foreign MPs require a fine-tuning of their expression to minimize the toxicity associated with their production. Here we report a novel regulation mechanism for the T7 expression system. We have isolated two bacterial hosts, namely C44(DE3) and C45(DE3), harboring a stop codon in the T7RNAP gene, whose translation is under the control of the basal nonsense suppressive activity of the BL21(DE3) host. Evaluation of hosts with superfolder green fluorescent protein (sfGFP) revealed an unprecedented tighter control of transgene expression with a marked accumulation of the recombinant protein during stationary phase. Analysis of a collection of twenty MP fused to GFP showed an improved production yield and quality of several bacterial MPs and of one human monotopic MP. These mutant hosts are complementary to the other existing T7 hosts and will increase the versatility of the T7 expression system.
- 45Ichiyanagi, K.; Beauregard, A.; Lawrence, S.; Smith, D.; Cousineau, B.; Belfort, M. Retrotransposition of the Ll.LtrB group II intron proceeds predominantly via reverse splicing into DNA targets. Mol. Microbiol. 2002, 46, 1259– 1272, DOI: 10.1046/j.1365-2958.2002.03226.xGoogle Scholar45https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XpsFGisL0%253D&md5=a588926c206e0bf19be2da6bb66a36c9Retrotransposition of the LI.LtrB group II intron proceeds predominantly via reverse splicing into DNA targetsIchiyanagi, Kenji; Beauregard, Arthur; Lawrence, Stacey; Smith, Dorie; Cousineau, Benoit; Belfort, MarleneMolecular Microbiology (2002), 46 (5), 1259-1272CODEN: MOMIEE; ISSN:0950-382X. (Blackwell Science Ltd.)Catalytic group II introns are mobile retroelements that invade cognate intronless genes via retrohoming, where the introns reverse splice into double-stranded DNA (dsDNA) targets. They can also retrotranspose to ectopic sites at low frequencies. Whereas our previous studies with a bacterial intron, LI.LtrB, supported frequent use of RNA targets during retrotransposition, recent expts. with a retrotransposition indicator gene indicate that DNA, rather than RNA, is a prominent target, with both dsDNA and single-stranded DNA (ssDNA) as possibilities. Thus retrotransposition occurs in both transcriptional sense and antisense orientations of target genes, and is largely independent of homologous DNA recombination and of the endonuclease function of the intron-encoded protein, LtrA. Models based on both dsDNA and ssDNA targeting are presented. Interestingly, retrotransposition is biased toward the template for lagging-strand DNA synthesis, which suggests the possibility of the replication folk as a source of ssDNA. Consistent with some use of ssDNA targets, many retrotransposition sites lack nucleotides crit. for the unwinding of target duplex DNA. Moreover, in vitro the intron reverse spliced into ssDNA more efficiently than dsDNA substrates for some of the retrotransposition sites. Furthermore, many bacterial group II introns reside on the lagging-strand template, hinting at a role for DNA replication in intron dispersal in nature.
- 46Dickson, L.; Huang, H. R.; Liu, L.; Matsuura, M.; Lambowitz, A. M.; Perlman, P. S. Retrotransposition of a yeast group II intron occurs by reverse splicing directly into ectopic DNA sites. Proc. Natl. Acad. Sci. U.S.A. 2001, 98, 13207– 13212, DOI: 10.1073/pnas.231494498Google Scholar46https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXosFygtrs%253D&md5=5bd9e382b0d94d8b6bb91f15232e59ccRetrotransposition of a yeast group II intron occurs by reverse splicing directly into ectopic DNA sitesDickson, Lorna; Huang, Hon-Ren; Liu, Lu; Matsuura, Manabu; Lambowitz, Alan M.; Perlman, Philip S.Proceedings of the National Academy of Sciences of the United States of America (2001), 98 (23), 13207-13212CODEN: PNASA6; ISSN:0027-8424. (National Academy of Sciences)Group II introns, the presumed ancestors of nuclear pre-mRNA introns, are site-specific retroelements. In addn. to "homing" to unoccupied sites in intronless alleles, group II introns transpose at low frequency to ectopic sites that resemble the normal homing site. Two general mechanisms have been proposed for group II intron transposition, one involving reverse splicing of the intron RNA directly into an ectopic DNA site, and the other involving reverse splicing into a site in RNA followed by reverse transcription and integration of the resulting cDNA by homologous recombination. Here, by using an "inverted-site" strategy, we show that the yeast mtDNA group II intron aI1 retrotransposes by reverse splicing directly into an ectopic DNA site. This same mechanism could account for other previously described ectopic transposition events in fungi and bacteria and may have contributed to the dispersal of group II introns into different genes.
- 47Coros, C. J.; Landthaler, M.; Piazza, C. L.; Beauregard, A.; Esposito, D.; Perutka, J.; Lambowitz, A. M.; Belfort, M. Retrotransposition strategies of the Lactococcus lactis Ll.LtrB group II intron are dictated by host identity and cellular environment. Mol. Microbiol. 2005, 56, 509– 524, DOI: 10.1111/j.1365-2958.2005.04554.xGoogle Scholar47https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXjsFWmtLs%253D&md5=d68f89a37bc9cd4bd4c9c8f403dd2fd5Retrotransposition strategies of the lactococcus lactis Ll.LtrB group II intron are dictated by host identity and cellular environmentCoros, Colin J.; Landthaler, Markus; Piazza, Carol Lyn; Beauregard, Arthur; Esposito, Donna; Perutka, Jiri; Lambowitz, Alan M.; Belfort, MarleneMolecular Microbiology (2005), 56 (2), 509-524CODEN: MOMIEE; ISSN:0950-382X. (Blackwell Publishing Ltd.)Group II introns are mobile retroelements that invade their cognate intron-minus gene in a process known as retrohoming. They can also retrotranspose to ectopic sites at low frequency. Previous studies of the Lactococcus lactis intron Ll.LtrB indicated that in its native host, as in Escherichia coli, retrohoming occurs by the intron RNA reverse splicing into double-stranded DNA (dsDNA) through an endonuclease-dependent pathway. However, in retrotransposition in L. lactis, the intron inserts predominantly into single-stranded DNA (ssDNA), in an endonuclease-independent manner. This work describes the retrotransposition of the Ll.LtrB intron in E. coli, using a retrotransposition indicator gene previously employed in our L. lactis studies. Unlike in L. lactis, in E. coli, Ll.LtrB retrotransposed frequently into dsDNA, and the process was dependent on the endonuclease activity of the intron-encoded protein. Further, the endonuclease-dependent insertions preferentially occurred around the origin and terminus of chromosomal DNA replication. Insertions in E. coli can also occur through an endonuclease-independent pathway, and, as in L. lactis, such events have a more random integration pattern. Together these findings show that Ll.LtrB can retrotranspose through at least two distinct mechanisms and that the host environment influences the choice of integration pathway. Addnl., growth conditions affect the insertion pattern. We propose a model in which DNA replication, compactness of the nucleoid and chromosomal localization influence target site preference.
- 48Flynn, P. J.; Reece, R. J. Activation of Transcription by Metabolic Intermediates of the Pyrimidine Biosynthetic Pathway. Mol. Cell. Biol. 1999, 19, 882– 888, DOI: 10.1128/MCB.19.1.882Google Scholar48https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1MXhvFyitA%253D%253D&md5=c3a936b1eda8f038270317fa526b055eActivation of transcription by metabolic intermediates of the pyrimidine biosynthetic pathwayFlynn, Paul J.; Reece, Richard J.Molecular and Cellular Biology (1999), 19 (1), 882-888CODEN: MCEBD4; ISSN:0270-7306. (American Society for Microbiology)Saccharomyces cerevisiae responds to pyrimidine starvation by increasing the expression of four URA genes, encoding the enzymes of de novo pyrimidine biosynthesis, three- to eightfold. The increase in gene expression is dependent on a transcriptional activator protein, Ppr1p. Here, we investigate the mechanism by which the transcriptional activity of Ppr1p responds to the level of pyrimidine biosynthetic intermediates. We find that purified Ppr1p is unable to promote activation of transcription in an in vitro system. Transcriptional activation by Ppr1p can be obsd., however, if either dihydroorotic acid (DHO) or orotic acid (OA) is included in the transcription reactions. The transcriptional activation function and the DHO/OA-responsive element of Ppr1p localize to the carboxyl-terminal 134 amino acids of the protein. Thus, Ppr1p directly senses the level of early pyrimidine biosynthetic intermediates within the cell and activates the expression of genes encoding proteins required later in the pathway. These results are discussed in terms of (i) regulation of the pyrimidine biosynthetic pathway and (ii) a novel mechanism of regulating gene expression.
- 49Boeke, J. D.; La Croute, F.; Fink, G. R. A positive selection for mutants lacking orotidine-5′-phosphate decarboxylase activity in yeast: 5-fluoro-orotic acid resistance. Mol. Gen. Genet. 1984, 197, 345– 346, DOI: 10.1007/BF00330984Google Scholar49https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL2MXhs1elsrg%253D&md5=1fbe754fa41c9f426317388cbc7381abA positive selection for mutants lacking orotidine-5'-phosphate decarboxylase activity in yeast: 5-fluoroorotic acid resistanceBoeke, Jef D.; LaCroute, Francois; Fink, Gerald R.Molecular and General Genetics (1984), 197 (2), 345-6CODEN: MGGEAE; ISSN:0026-8925.Mutations at the URA3 locus of Saccharomyces cerevisiae can be obtained by a pos. selection. Wild-type strains of yeast (or ura3 mutant strains contg. a plasmid-borne URA3+ gene) are unable to grow on medium contg. the pyrimidine analog 5-fluoroorotic acid, whereas ura3- mutants grow normally. This selection, based on the loss of orotidine-5'-phosphate decarboxylase activity seems applicable to a variety of eucaryotic and procaryotic cells.
- 50Galvão, T. C.; De Lorenzo, V. Adaptation of the yeast URA3 selection system to Gram-negative bacteria and generation of a ΔbetCDE Pseudomonas putida strain. Appl. Environ. Microbiol. 2005, 71, 883– 892, DOI: 10.1128/AEM.71.2.883-892.2005Google Scholar50https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2MXhsVKgtrw%253D&md5=03f764b6e6c5e72eb582e0cfcf77882bAdaptation of the yeast URA3 selection system to Gram-negative bacteria and generation of a ΔbetCDE Pseudomonas putida strainGalvao, Teca Calcagno; De Lorenzo, VictorApplied and Environmental Microbiology (2005), 71 (2), 883-892CODEN: AEMIDF; ISSN:0099-2240. (American Society for Microbiology)A general procedure for efficient generation of gene knockouts in gram-neg. bacteria by the adaptation of the Saccharomyces cerevisiae URA3 selection system is described. A Pseudomonas putida strain lacking the URA3 homolog pyrF (encoding orotidine-5'-phosphate decarboxylase) was constructed, allowing the use of a plasmid-borne copy of the gene as the target of selection. The delivery vector pTEC contains the pyrF gene and promoter, a conditional origin of replication (oriR6K), an origin of transfer (mobRK2), and an antibiotic selection marker flanked by multiple sites for cloning appropriate DNA segments. The versatility of pyrF as a selection system, allowing both pos. and neg. selection of the marker, and the robustness of the selection, where pyrF is assocd. with uracil prototrophy and fluoroorotic acid sensitivity, make this setup a powerful tool for efficient homologous gene replacement in gram-neg. bacteria. The system has been instrumental for complete deletion of the P. putida choline-O-sulfate utilization operon betCDE, a mutant which could not be produced by any of the other genetic strategies available.
- 51Cousineau, B.; Lawrence, S.; Smith, D.; Belfort, M. Retrotransposition of a bacterial group II intron. Nature 2000, 404, 1018– 1021, DOI: 10.1038/35010029Google Scholar51https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXjtFynu7s%253D&md5=8ce276b86062da0e7ec9423eae506dd0Retrotransposition of a bacterial group II intronCousineau, Benoit; Lawrence, Stacey; Smith, Dorie; Belfort, MarleneNature (London) (2000), 404 (6781), 1018-1021CODEN: NATUAS; ISSN:0028-0836. (Nature Publishing Group)Self-splicing group II introns may be the evolutionary progenitors of eukaryotic spliceosomal introns, but the route by which they invade new chromosomal sites is unknown. To address the mechanism by which group II introns are disseminated, we have studied the bacterial L1.LtrB intron from Lactococcus lactis. The protein product of this intron, LtrA, possesses maturase, reverse transcriptase and endonuclease enzymic activities. Together with the intron, LtrA forms a ribonucleoprotein (RNP) complex which mediates a process known as retrohoming. In retrohoming, the intron reverse splices into a cognate intronless DNA site. Integration of a DNA copy of the intron is recombinase independent but requires all three activities of LtrA. Here we report the first exptl. demonstration of a group II intron invading ectopic chromosomal sites, which occurs by a distinct retrotransposition mechanism. This retrotransposition process is endonuclease-independent and recombinase-dependent, and is likely to involve reverse splicing of the intron RNA into cellular RNA targets. These retrotranspositions suggest a mechanism by which splicesomal introns may have become widely dispersed.
- 52Golomb, M.; Chamberlin, M. Characterization of T7 specific ribonucleic acid polymerase. IV. Resolution of the major in vitro transcripts by gel electrophoresis. J. Biol. Chem. 1974, 249, 2858– 2863, DOI: 10.1016/S0021-9258(19)42709-9Google Scholar52https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaE2cXksVKrs70%253D&md5=5abe8d2ccb186b1d414061a838117d69Characterization of T7-specific ribonucleic acid polymerase. IV. Resolution of the major in vitro transcripts by gel electrophoresisGolomb, Miriam; Chamberlin, MichaelJournal of Biological Chemistry (1974), 249 (9), 2858-63CODEN: JBCHA3; ISSN:0021-9258.The major in vitro RNA transcripts synthesized by T7 RNA polymerase with T7 and T3 DNA templates were resolved by electrophoresis on polyacrylamide gels. Six discrete size classes of T7 RNAs were found, and designated I to VI. The apparent mol. wts., estd. from their electrophoretic mobilities, were 2 × 105-5 × 106. At least 2 minor RNA species with mol. wts. >5 × 106 were also detected. The 6 major T7 RNA species were synthesized in approx. equimolar amts., with the exception of species III, which was made in ∼ twice this amt. Perhaps, species III is a mixt. of two RNAs transcribed from sep. regions of the T7 genome. Hence, the 6 major T7 RNA species are tentatively identified with 7 late transcription units on the phage chromosome, which are read with equal efficiencies. The 6 (or 7) transcription products were inititated independently with GTP at the 5' terminus, and were elongated at a rate of 230 nucleotides/s under standard in vitro conditions. When T7 RNA polymerase was used to transcribe T3 DNA, a single RNA transcript was found, with a mol. wt. similar to that of T7 RNA species III. This suggests an explanation for the reduced rate of T3 RNA synthesis by T7 RNA polymerase in vitro, and implies that T3 promoter sites read by the T3 RNA polymerase are heterogeneous in nature.
- 53Chamberlin, M.; Ryan, T. 4 Bacteriophage DNA-Dependent RNA Polymerases. Enzymes 1982, 15, 87– 108Google Scholar53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL38XltVGqsLo%253D&md5=edd42f3f5bbcbd40064d5fdb469be1e4Bacteriophage DNA-dependent RNA polymerasesChamberlin, Michael J.; Ryan, T.(1982), 15 (Pt. B), 87-108CODEN: 25GLAS ISSN:. (Academic)A review with 87 refs.
- 54Benedetti, I.; de Lorenzo, V.; Nikel, P. I. Genetic programming of catalytic Pseudomonas putida biofilms for boosting biodegradation of haloalkanes. Metab. Eng. 2016, 33, 109– 118, DOI: 10.1016/j.ymben.2015.11.004Google Scholar54https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXhvFems7rL&md5=64291830c898a5f4ecfb5fef5313598cGenetic programming of catalytic Pseudomonas putida biofilms for boosting biodegradation of haloalkanesBenedetti, Ilaria; de Lorenzo, Victor; Nikel, Pablo I.Metabolic Engineering (2016), 33 (), 109-118CODEN: MEENFM; ISSN:1096-7176. (Elsevier B. V.)Bacterial biofilms outperform planktonic counterparts in whole-cell biocatalysis. The transition between planktonic and biofilm lifestyles of the platform strain Pseudomonas putida KT2440 is ruled by a regulatory network controlling the levels of the trigger signal cyclic di-GMP (c-di-GMP). This circumstance was exploited for designing a genetic device that over-runs the synthesis or degrdn. of c-di-GMP, thus making P. putida to form biofilms at user's will. For this purpose, the transcription of either yedQ (diguanylate cyclase) or yhjH (c-di-GMP phosphodiesterase) from Escherichia coli was artificially placed under the tight control of a cyclohexanone-responsive expression system. The resulting strain was subsequently endowed with a synthetic operon and tested for 1-chlorobutane biodegrdn. Upon addn. of cyclohexanone to the culture medium, the thereby designed P. putida cells formed biofilms displaying high dehalogenase activity. These results show that the morphologies and phys. forms of whole-cell biocatalysts can be genetically programmed while purposely designing their biochem. activity.
- 55Martínez-Abarca, F.; García-Rodríguez, F. M.; Toro, N. Homing of a bacterial group II intron with an intron-encoded protein lacking a recognizable endonuclease domain. Mol. Microbiol. 2000, 35, 1405– 1412, DOI: 10.1046/j.1365-2958.2000.01804.xGoogle Scholar55https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXis1Wrtro%253D&md5=7cf3687672ca684154f02972891ab74bHoming of a bacterial group II intron with an intron-encoded protein lacking a recognizable endonuclease domainMartinez-Abarca, Francisco; Garcia-Rodriguez, Fernando M.; Toro, NicolasMolecular Microbiology (2000), 35 (6), 1405-1412CODEN: MOMIEE; ISSN:0950-382X. (Blackwell Science Ltd.)Rmlnt1 is a functional group II intron found in Sinorhizobium meliloti where it interrupts a group of IS elements of the IS630-Tc1 family. In contrast to many other group II introns, the intron-encoded protein (IEP) of Rmlnt1 lacks the characteristic conserved part of the Zn domain assocd. with the IEP endonuclease activity. Nevertheless, in this study, we show that Rmlnt1 is capable of inserting into a vector contg. the DNA spanning the Rmlnt1 target site from the genome of S. meliloti. Efficient homing was also obsd. in the absence of homologous recombination (RecA- strains). In addn., it is shown that Rmlnt1 is able to move to its target in a heterologous host (S. medicae). Homing of Rmlnt1 occurs very efficiently upon DNA target uptake (conjugation/electroporation) by the host cell resulting in a proportion of invaded target of 11-30%. Afterwards, the remaining intronless target DNA is protected from intron invasion.
- 56Cook, T. B.; Rand, J. M.; Nurani, W.; Courtney, D. K.; Liu, S. A.; Pfleger, B. F. Genetic tools for reliable gene expression and recombineering in Pseudomonas putida. J. Ind. Microbiol. Biotechnol. 2018, 45, 517– 527, DOI: 10.1007/s10295-017-2001-5Google Scholar56https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXjt1GltA%253D%253D&md5=8404ada292d1376d0fa2b08c54d1863aGenetic tools for reliable gene expression and recombineering in Pseudomonas putidaCook, Taylor B.; Rand, Jacqueline M.; Nurani, Wasti; Courtney, Dylan K.; Liu, Sophia A.; Pfleger, Brian F.Journal of Industrial Microbiology & Biotechnology (2018), 45 (7), 517-527CODEN: JIMBFL; ISSN:1367-5435. (Springer)Pseudomonas putida is a promising bacterial host for producing natural products, such as polyketides and nonribosomal peptides. In these types of projects, researchers need a genetic toolbox consisting of plasmids, characterized promoters, and techniques for rapidly editing the genome. Past reports described constitutive promoter libraries, a suite of broad host range plasmids that replicate in P. putida, and genome-editing methods. To augment those tools, we have characterized a set of inducible promoters and discovered that IPTG-inducible promoter systems have poor dynamic range due to overexpression of the LacI repressor. By replacing the promoter driving lacI expression with weaker promoters, we increased the fold induction of an IPTG-inducible promoter in P. putida KT2440 to 80-fold. Upon discovering that gene expression from a plasmid was unpredictable when using a high-copy mutant of the BBR1 origin, we detd. the copy nos. of several broad host range origins and found that plasmid copy nos. are significantly higher in P. putida KT2440 than in the synthetic biol. workhorse, Escherichia coli. Lastly, we developed a λRed/Cas9 recombineering method in P. putida KT2440 using the genetic tools that we characterized. This method enabled the creation of scarless mutations without the need for performing classic two-step integration and marker removal protocols that depend on selection and counterselection genes. With the method, we generated four scarless deletions, three of which we were unable to create using a previously established genome-editing technique.
- 57Bagdasarian, M.; Lurz, R.; Rückert, B.; Franklin, F. C. H.; Bagdasarian, M. M.; Frey, J.; Timmis, K. N. Specific-purpose plasmid cloning vectors II. Broad host range, high copy number, RSF 1010-derived vectors, and a host-vector system for gene cloning in Pseudomonas. Gene 1981, 16, 237– 247, DOI: 10.1016/0378-1119(81)90080-9Google Scholar57https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL38XhsFKlsrg%253D&md5=886ab175b0c338a0ec3030fe6e38c178Specific-purpose plasmid cloning vectors. II. Broad host range, high copy number, RSF1010-derived vectors, and a host-vector system for gene cloning in PseudomonasBagdasarian, M.; Lurz, R.; Rueckert, B.; Franklin, F. C. H.; Bagdasarian, M. M.; Frey, J.; Timmis, K. N.Gene (1981), 16 (1-3), 237-47CODEN: GENED6; ISSN:0378-1119.Host-vector systems were developed for gene cloning in the metabolically versatile bacterial genus Pseudomonas. They comprise restriction-neg. host strains of P. aeruginosa and P. putida and new cloning vectors derived from the high-copy-no., broad-host-range plasmid RSF1010, which are stably maintained in a wide range of gram-neg. bacteria. These plasmids contain EcoRI, SstI, HindIII, XmaI, XhoI, SalI, BamHI, and ClaI insertion sites. All cloning sites, except for BamHI and ClaI, are located within antibiotic-resistance genes; insertional inactivation of these genes during hybrid plasmid formation provides a readily scored phenotypic change for the rapid identification of bacterial clones carrying such hybrids. One of the new vector plasmids is a cosmid that may be used for the selective cloning of large DNA fragments by in vitro λ packaging. An analogous series of vectors that are defective in their plasmid-mobilization function, and that exhibit a degree of biol. containment comparable to that of current Escherichia coli vector plasmids, are also described.
- 58Jahn, M.; Vorpahl, C.; Hübschmann, T.; Harms, H.; Müller, S. Copy number variability of expression plasmids determined by cell sorting and Droplet Digital PCR. Microb. Cell Fact. 2016, 15, 211 DOI: 10.1186/s12934-016-0610-8Google Scholar58https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1cXjsFCrsbc%253D&md5=3245e40449919390aec4633a06832341Copy number variability of expression plasmids determined by cell sorting and Droplet Digital PCRJahn, Michael; Vorpahl, Carsten; Huebschmann, Thomas; Harms, Hauke; Mueller, SusannMicrobial Cell Factories (2016), 15 (), 211/1-211/12CODEN: MCFICT; ISSN:1475-2859. (BioMed Central Ltd.)Plasmids are widely used for mol. cloning or prodn. of proteins in lab. and industrial settings. Const. modification has brought forth countless plasmid vectors whose characteristics in terms of av. plasmid copy no. (PCN) and stability are rarely known. The crucial factor detg. the PCN is the replication system; most replication systems in use today belong to a small no. of different classes and are available through repositories like the Std. European Vector Architecture (SEVA). In this study, the PCN was detd. in a set of seven SEVA-based expression plasmids only differing in the replication system. The av. PCN for all constructs was detd. by Droplet Digital PCR and ranged between 2 and 40 per chromosome in the host organism Escherichia coli. Furthermore, a plasmid-encoded EGFP reporter protein served as a means to assess variability in reporter gene expression on the single cell level. Only cells with one type of plasmid (RSF1010 replication system) showed a high degree of heterogeneity with a clear bimodal distribution of EGFP intensity while the others showed a normal distribution. The heterogeneous RSF1010-carrying cell population and one normally distributed population (ColE1 replication system) were further analyzed by sorting cells of sub-populations selected according to EGFP intensity. For both plasmids, low and highly fluorescent sub-populations showed a remarkable difference in PCN, ranging from 9.2 to 123.4 for ColE1 and from 0.5 to 11.8 for RSF1010, resp. The av. PCN detd. here for a set of standardized plasmids was generally at the lower end of previously reported ranges and not related to the degree of heterogeneity. Further characterization of a heterogeneous and a homogeneous population demonstrated considerable differences in the PCN of sub-populations. We therefore present direct mol. evidence that the av. PCN does not represent the true no. of plasmid mols. in individual cells.
- 59Pyne, M. E.; Moo-Young, M.; Chung, D. A.; Chou, C. P. Coupling the CRISPR/Cas9 system with lambda red recombineering enables simplified chromosomal gene replacement in Escherichia coli. Appl. Environ. Microbiol. 2015, 81, 5103– 5114, DOI: 10.1128/AEM.01248-15Google Scholar59https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXht1Ojt7bP&md5=993a4d147fcb4c66368e6ec73bae3f7aCoupling the CRISPR/Cas9 system with lambda Red recombineering enables simplified chromosomal gene replacement in Escherichia coliPyne, Michael E.; Moo-Young, Murray; Chung, Duane A.; Chou, C. PerryApplied and Environmental Microbiology (2015), 81 (15), 5103-5114CODEN: AEMIDF; ISSN:1098-5336. (American Society for Microbiology)To date, most genetic engineering approaches coupling the type II Streptococcus pyogenes clustered regularly interspaced short palindromic repeat (CRISPR)/Cas9 system to lambda Red recombineering have involved minor single nucleotide mutations. Here we show that procedures for carrying out more complex chromosomal gene replacements in Escherichia coli can be substantially enhanced through implementation of CRISPR/Cas9 genome editing. We developed a three-plasmid approach that allows not only highly efficient recombination of short single-stranded oligonucleotides but also replacement of multigene chromosomal stretches of DNA with large PCR products. By systematically challenging the proposed system with respect to the magnitude of chromosomal deletion and size of DNA insertion, we demonstrated DNA deletions of up to 19.4 kb, encompassing 19 nonessential chromosomal genes, and insertion of up to 3 kb of heterologous DNA with recombination efficiencies permitting mutant detection by colony PCR screening. Since CRISPR/Cas9-coupled recombineering does not rely on the use of chromosome- encoded antibiotic resistance, or flippase recombination for antibiotic marker recycling, our approach is simpler, less labor-intensive, and allows efficient prodn. of gene replacement mutants that are both markerless and "scar"-less.
- 60Moreb, E. A.; Hoover, B.; Yaseen, A.; Valyasevi, N.; Roecker, Z.; Menacho-Melgar, R.; Lynch, M. D. Managing the SOS Response for Enhanced CRISPR-Cas-Based Recombineering in E. coli through Transient Inhibition of Host RecA Activity. ACS Synth. Biol. 2017, 6, 2209– 2218, DOI: 10.1021/acssynbio.7b00174Google Scholar60https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhsV2murrN&md5=70fcee7818b69824c249a54d3385fe39Managing the SOS Response for Enhanced CRISPR-Cas-Based Recombineering in E. coli through Transient Inhibition of Host RecA ActivityMoreb, Eirik Adim; Hoover, Benjamin; Yaseen, Adam; Valyasevi, Nisakorn; Roecker, Zoe; Menacho-Melgar, Romel; Lynch, Michael D.ACS Synthetic Biology (2017), 6 (12), 2209-2218CODEN: ASBCD6; ISSN:2161-5063. (American Chemical Society)Phage-derived "recombineering" methods are utilized for bacterial genome editing. Recombineering results in a heterogeneous population of modified and unmodified chromosomes, and therefore selection methods, such as CRISPR-Cas9, are required to select for edited clones. Cells can evade CRISPR-Cas-induced cell death through recA-mediated induction of the SOS response. The SOS response increases RecA dependent repair as well as mutation rates through induction of the umuDC error prone polymerase. As a result, CRISPR-Cas selection is more efficient in recA mutants. We report an approach to inhibiting the SOS response and RecA activity through the expression of a mutant dominant neg. form of RecA, which incorporates into wild type RecA filaments and inhibits activity. Using a plasmid-based system in which Cas9 and recA mutants are coexpressed, we can achieve increased efficiency and consistency of CRISPR-Cas9-mediated selection and recombineering in E. coli, while reducing the induction of the SOS response. To date, this approach has been shown to be independent of recA genotype and host strain lineage. Using this system, we demonstrate increased CRISPR-Cas selection efficacy with over 10,000 guides covering the E. coli chromosome. The use of dominant neg. RecA or homologues may be of broad use in bacterial CRISPR-Cas-based genome editing where the SOS pathways are present.
- 61Tellechea-Luzardo, J.; Winterhalter, C.; Widera, P.; Kozyra, J.; De Lorenzo, V.; Krasnogor, N. Linking Engineered Cells to Their Digital Twins: A Version Control System for Strain Engineering. ACS Synth. Biol. 2020, 9, 536– 545, DOI: 10.1021/acssynbio.9b00400Google Scholar61https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3cXjsVOntr0%253D&md5=282b32256dcc6a167abd1fb5b8484f29Linking Engineered Cells to Their Digital Twins: A Version Control System for Strain EngineeringTellechea-Luzardo, Jonathan; Winterhalter, Charles; Widera, Pawel; Kozyra, Jerzy; de Lorenzo, Victor; Krasnogor, NatalioACS Synthetic Biology (2020), 9 (3), 536-545CODEN: ASBCD6; ISSN:2161-5063. (American Chemical Society)As DNA sequencing and synthesis become cheaper and more easily accessible, the scale and complexity of biol. engineering projects is set to grow. Yet, although there is an accelerating convergence between biotechnol. and digital technol., a deficit in software and lab. techniques diminishes the ability to make biotechnol. more agile, reproducible and transparent while, at the same time, limiting the security and safety of synthetic biol. constructs. To partially address some of these problems, this paper presents an approach for phys. linking engineered cells to their digital footprint-we called it digital twinning. This enables the tracking of the entire engineering history of a cell line in a specialized version control system for collaborative strain engineering via simple barcoding protocols.
- 62Tellechea-Luzardo, J.; Hobbs, L.; Velázquez, E.; Pelechova, L.; Woods, S.; de Lorenzo, V.; Krasnogor, N. Versioning Biological Cells for Trustworthy Cell Engineering. bioRxiv 2021. DOI: 10.1101/2021.04.23.441106 .Google ScholarThere is no corresponding record for this reference.
- 63Zimmerly, S.; Hausner, G.; Wu, X. C. Phylogenetic relationships among group II intron ORFs. Nucleic Acids Res. 2001, 29, 1238– 1250, DOI: 10.1093/nar/29.5.1238Google Scholar63https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXhvFGgsbk%253D&md5=a70b635e53fde84e69afc1f5b25a7d4fPhylogenetic relationships among group II intron ORFsZimmerly, Steven; Hausner, Georg; Wu, Xu-ChuNucleic Acids Research (2001), 29 (5), 1238-1250CODEN: NARHAD; ISSN:0305-1048. (Oxford University Press)Group II introns are widely believed to have been ancestors of spliceosomal introns, yet little is known about their own evolutionary history. In order to address the evolution of mobile group II introns, the authors have compiled 71 open reading frames (ORFs) related to group II intron reverse transcriptases and subjected their derived amino acid sequences to phylogenetic anal. The phylogenetic tree was rooted with reverse transcriptases (RTs) of non-long terminal repeat retroelements, and the inferred phylogeny reveals two major clusters which the authors term the mitochondrial and chloroplast-like lineages. Bacterial ORFs are mainly positioned at the bases of the two lineages but with weak bootstrap support. The data give an overview of an apparently high degree of horizontal transfer of group II intron ORFs, mostly among related organisms but also between organelles and. bacteria. The Zn domain (nuclease) and YADD motif (RT active site) were lost multiple times during evolution. Differences in domain structures suggest that the oldest ORFs were concise, while the ORF in the mitochondrial lineage subsequently expanded in three locations. The data are consistent with a bacterial origin for mobile group II introns.
- 64More, S. J.; Bampidis, V.; Benford, D.; Bragard, C.; Halldorsson, T. I.; Hernández-Jerez, A.; Susanne, H. B.; Koutsoumanis, K.; Machera, K.; Naegeli, H.; Nielsen, S. S.; Schlatter, J.; Schrenk, D.; Silano, V.; Turck, D.; Younes, M.; Glandorf, B.; Herman, L.; Tebbe, C.; Vlak, J.; Aguilera, J.; Schoonjans, R.; Cocconcelli, P. S. Evaluation of existing guidelines for their adequacy for the microbial characterisation and environmental risk assessment of microorganisms obtained through synthetic biology. EFSA J. 2020, 18, 1– 50Google ScholarThere is no corresponding record for this reference.
- 65Horton, R. M.; Hunt, H. D.; Ho, S. N.; Pullen, J. K.; Pease, L. R. Engineering hybrid genes without the use of restriction enzymes: gene splicing by overlap extension. Gene 1989, 77, 61– 68, DOI: 10.1016/0378-1119(89)90359-4Google Scholar65https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL1MXktVais7k%253D&md5=40edd5de9a603f20bb901e78a3e3588dEngineering hybrid genes without the use of restriction enzymes: gene splicing by overlap extensionHorton, Robert M.; Hunt, Henry D.; Ho, Steffan N.; Pullen, Jeffrey K.; Pease, Larry R.Gene (1989), 77 (1), 61-8CODEN: GENED6; ISSN:0378-1119.Gene splicing by overlap extension is a new approach for recombining DNA mols. at precise junctions irresp. of nucleotide sequences at the recombination site and without the use of restriction endonucleases or ligase. Fragments from the genes that are to be recombined are generated in sep. polymerase chain reactions (PCRs). The primers are designed so that the ends of the products contain complementary sequences. When these PCR products are mixed, denatured, and reannealed, the strands having the matching sequences at their 3' ends overlap and act as primers for each other. Extension of this overlap by DNA polymerase produces a mol. in which the original sequences are spliced together. This technique is used to construct a gene encoding a mosaic fusion protein comprised of parts of two different mouse class-I major histocompatibility genes. This simple and widely applicable approach has significant advantages over std. recombinant DNA techniques.
- 66Gibson, D. G.; Young, L.; Chuang, R. Y.; Venter, J. C.; Hutchison, C. A.; Smith, H. O. Enzymatic assembly of DNA molecules up to several hundred kilobases. Nat. Methods 2009, 6, 343– 345, DOI: 10.1038/nmeth.1318Google Scholar66https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXksVemsbw%253D&md5=46284924c7d73c47cfb490983338e480Enzymatic assembly of DNA molecules up to several hundred kilobasesGibson, Daniel G.; Young, Lei; Chuang, Ray-Yuan; Venter, J. Craig; Hutchison, Clyde A.; Smith, Hamilton O.Nature Methods (2009), 6 (5), 343-345CODEN: NMAEA3; ISSN:1548-7091. (Nature Publishing Group)The authors describe an isothermal, single-reaction method for assembling multiple overlapping DNA mols. by the concerted action of a 5' exonuclease, a DNA polymerase and a DNA ligase. First they recessed DNA fragments, yielding single-stranded DNA overhangs that specifically annealed, and then covalently joined them. This assembly method can be used to seamlessly construct synthetic and natural genes, genetic pathways and entire genomes, and could be a useful mol. engineering tool.
- 67Gibson, D. G.; Glass, J. I.; Lartigue, C.; Noskov, V. N.; Chuang, R. Y.; Algire, M. A.; Benders, G. A.; Montague, M. G.; Ma, L.; Moodie, M. M.; Merryman, C.; Vashee, S.; Krishnakumar, R.; Assad-Garcia, N.; Andrews-Pfannkoch, C.; Denisova, E. A.; Young, L.; Qi, Z. N.; Segall-Shapiro, T. H.; Calvey, C. H.; Parmar, P. P.; Hutchison, C. A.; Smith, H. O.; Venter, J. C. Creation of a bacterial cell controlled by a chemically synthesized genome. Science 2010, 329, 52– 56, DOI: 10.1126/science.1190719Google Scholar67https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXotVeqsLg%253D&md5=f978c98f3e5c3c9b14bd4c7d8a1eecc7Creation of a bacterial cell controlled by a chemically synthesized genomeGibson, Daniel G.; Glass, John I.; Lartigue, Carole; Noskov, Vladimir N.; Chuang, Ray-Yuan; Algire, Mikkel A.; Benders, Gwynedd A.; Montague, Michael G.; Ma, Li; Moodie, Monzia M.; Merryman, Chuck; Vashee, Sanjay; Krishnakumar, Radha; Assad-Garcia, Nacyra; Andrews-Pfannkoch, Cynthia; Denisova, Evgeniya A.; Young, Lei; Qi, Zhi-Qing; Segall-Shapiro, Thomas H.; Calvey, Christopher H.; Parmar, Prashanth P.; Hutchison, Clyde A., III; Smith, Hamilton O.; Venter, J. CraigScience (Washington, DC, United States) (2010), 329 (5987), 52-56CODEN: SCIEAS; ISSN:0036-8075. (American Association for the Advancement of Science)We report the design, synthesis, and assembly of the 1.08-mega-base pair Mycoplasma mycoides JCVI-syn1.0 genome starting from digitized genome sequence information and its transplantation into a M. capricolum recipient cell to create new M. mycoides cells that are controlled only by the synthetic chromosome. The only DNA in the cells is the designed synthetic DNA sequence, including "watermark" sequences and other designed gene deletions and polymorphisms, and mutations acquired during the building process. The new cells have expected phenotypic properties and are capable of continuous self-replication. The complete, annotated synthetic genome sequence is deposited in GenBank/EMBL/DDBJ with accession no. CP002027.
Supporting Information
Supporting Information
ARTICLE SECTIONSThe Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acssynbio.1c00199.
List of oligonucleotides used in this study (Supplementary Table S1); list of plasmids used in this work (Supplementary Table S2); insertion frequencies of Ll.LtrB::Lux1 and Ll.LtrB::Lux4 in P. putida KT2440 WT and ΔrecA with no CRISPR/Cas9-mediated counterselection (Supplementary Table S3); insertion frequency of Ll.LtrB::LuxN intron in P. putida KT2440 WT with 5FOA CRISPR/Cas9-mediated counterselection (Supplementary Table S4); insertion frequency of Ll.LtrB::LuxN intron in P. putida KT2440 ΔrecA with 5FOA CRISPR/Cas9-mediated counterselection (Supplementary Table S5); construction and verification of intron delivery plasmids compatible with CRISPR/Cas9-mediated counterselection (Supplementary Methods); pSEVA plasmids for the expression of the Ll.LtrB intron in a wide range of Gram-negative bacteria (Supplementary Figure S1); preliminary approach to assess size-restriction of intron-mediated delivery using luxC fragments as a cargo and 5FOA counterselection in log-phase induced cells (Supplementary Figure S2); total intron insertion frequency without differentiating the cargo that was being delivered (Supplementary Figure S3); barcode generation with a PCR using 3′-overlapping 119-mer oligonucleotides (Supplementary Figure S4); application of targetrons as a barcode delivery system (Supplementary Figure S5); and design and test of Locus 1 and 2 spacers for CRISPR/Cas9-mediated counterselection of Ll.LtrB::B3 group II intron (Supplementary Figure S6) (PDF)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.