Dokument: Establishing regulatable expression systems in the acetic acid bacterium Gluconobacter oxydans 621H

Titel:Establishing regulatable expression systems in the acetic acid bacterium Gluconobacter oxydans 621H
Weiterer Titel:Etablierung regulierbarer Expressionssysteme in dem Essigsäurebakterium Gluconobacter oxydans 621H
URL für Lesezeichen:https://docserv.uni-duesseldorf.de/servlets/DocumentServlet?id=58806
URN (NBN):urn:nbn:de:hbz:061-20220307-105908-0
Kollektion:Dissertationen
Sprache:Englisch
Dokumententyp:Wissenschaftliche Abschlussarbeiten » Dissertation
Medientyp:Text
Autor:Dr. Fricke, Philipp Moritz [Autor]
Dateien:
[Dateien anzeigen]Adobe PDF
[Details]34,08 MB in einer Datei
[ZIP-Datei erzeugen]
Dateien vom 10.02.2022 / geändert 10.02.2022
Beitragende:Prof. Dr. Bott, Michael [Gutachter]
Prof. Dr. Urlacher, Vlada [Gutachter]
Stichwörter:Bacteria, Acetic acid bacteria, Gluconobacter oxydans, Biotechnology, Genetics, Expression system, Fluorescence protein
Dewey Dezimal-Klassifikation:500 Naturwissenschaften und Mathematik » 570 Biowissenschaften; Biologie
Beschreibungen:Das Essigsäurebakterium Gluconobacter oxydans wird unter den industriell relevanten Mikroorganismen für seine Fähigkeit geschätzt, eine Vielzahl von Kohlenhydraten stereo- und regiospezifisch unvollständig zu oxidieren. Bisherige biotechnologische Verfahren mit G. oxydans verwendeten für die zielgerichtete Genexpression ausschließlich konstitutive Promotoren. Die regulierbaren Promotoren, die in G. oxydans oder anderen Essigsäurebakterien untersucht wurden, zeigten eine sehr hohe Basalaktivität unter nicht induzierten Bedingungen, was zu einem geringen Induktionsverhältnis führte. Das Ziel dieser Doktorarbeit war es, ein regulierbares Promotorsystem in G. oxydans zu etablieren, dass die Expression ausgewählter Gene durch die Konzentration eines Induktors steuern kann. Zu diesem Zweck wurden vier bekannte und gut charakterisierte Regulator-Promotor-Paare in G.oxydans mit Hilfe von Expressionsplasmiden getestet. Außerdem wurde versucht, regulierbare endogene G. oxydans-Promotoren zu identifizieren, die auf die Zugabe eines Metaboliten als externen Stimulus reagieren und so für eine kontrollierte Genexpression geeignet sind.
In G. oxydans getestet, ermöglichte das AraC-ParaBAD-System aus Escherichia coli MC4100 eine starke Repression der Transkription in Abwesenheit von L-Arabinose sowie eine bis zu 480-fach erhöhte Genexpression bei vollständiger Induktion mit L-Arabinose. Die Genexpression von ParaBAD war durch den Einsatz von Induktorkonzentrationen von 0,1-1% (w/v) L-Arabinose gut modulierbar. Außerdem wurde gezeigt, dass die Oxidation von L-Arabinose zu L-Arabinsäure durch die membrangebundene Glucose-Dehydrogenase GdhM zu einer Ansäuerung des Mediums und damit zu einem pH-abhängigen Verlust der Reporterproteinaktivität führte. In pH-kontrollierten Fermentationen oder durch die Verwendung des GdhM-Deletionsstamms BP.6 konnte dieser Verlust umgangen werden.
Das Oxidationsprodukt L-Arabinsäure ist ein wertvoller Rohstoff, der in vielen Prozessen unter anderem bei der Herstellung bestimmter Pharmazeutika, Halbleiter- materialien und Verbundzementen Anwendung findet. Durch die Verwendung von G. oxydans 621H wurde in pH-kontrollierten Fed-Batch-Kultivierungen innerhalb von 144 Stunden mehr als 120 g/L L-Arabinsäure aus 160 g/L L-Arabinose hergestellt. Dieser hohe Titer übertraf die Titer, die zuvor für gentechnisch veränderte E. coli oder Saccharomyces cerevisiae berichtet wurden, um mehr als das Dreifache. Die Plasmid-basierte Überexpression von gdhM zur Erhöhung der Enzymaktivität führte, möglicherweise aufgrund nachteiliger Auswirkungen auf die Membranintegrität, zu keiner Verbesserung der Produktivität.
Neben dem AraC-ParaBAD-System wurden mit dem Transposon Tn10-basiertem TetR- Ptet-System und dem LacI-PlacUV5-System zwei weitere induzierbare Expressionssysteme aus E. coli in G. oxydans getestet. Im Gegensatz zu AraC-ParaBAD beruhen diese induzierbaren Expressionssysteme auf der Transkriptionsrepression von Zielgenen. Die Leistung des TetR- Ptet-Systems übertraf die des AraC-ParaBAD-Systems. Unter nicht induzierten Bedingungen zeigte das TetR-Ptet-System eine sehr geringe basale Expression, die durch die Konzentration des Induktors graduell gesteigert werden konnte. Unter Verwendung von 200ng/ml Anhydrotetracyclin wurde ein maximales Induktionsverhältnis von 3.500 ermittelt. Terminatorsequenzen, Wahl des Expressionsplasmids, Antibiotikaresistenzgen und die Insertion einer zusätzlichen, bekannten Ribosomenbindestelle beeinflussten die Leistungsfähigkeit dieses Systems. Die deutlich geringere Induzierbarkeit von PlacUV5 in G. oxydans hingegen machte dieses System zum schwächsten der hier getesteten Regulator- Promotor-Paare. Aufgrund der hohen Basalaktivität wurde mit PlacUV5 nur eine 40-fache Induktion mit 1 mM IPTG erzielt.
Mit dem RhaSR-PrhaBAD-System von E. coli wurde ein Regulator-Promotor-Paar gefunden, das – im Gegensatz zu E. coli – einen L-Rhamnose-abhängigen Knockdown anstelle einer Induktion der Genexpression in G. oxydans ermöglichte. PrhaBAD führte in G. oxydans zu einer sehr starken Expression und wurde durch die Anwesenheit von RhaS in Abwesenheit von L-Rhamnose aktiviert. Mit steigender L-Rhamnose-Konzentration wurde die PrhaBAD-Aktivität RhaS-abhängig um bis zu 92% reduziert. Im Gegensatz dazu war die Expression des RhaS-abhängigen Permease-Promotors PrhaT in G. oxydans genauso wie in E. coli induzierbar. PrhaT zeigte in G. oxydans eine schwache Expressionsstärke, welche durch die L-Rhamnosekonzentration maximal 10-fach induziert werden konnte. Ein derartiger Promotor wäre möglicherweise für die Expression von „schwierigen“ Enzymen von Vorteil.
In DNA-Microarray-basierten genomweiten Stimulon-Screenings auf regulierte endogene Promotoren wurde gezeigt, dass sich die Transkription der G. oxydans-Gene GOX0532 und GOX0536 um das 45-fache erhöhte, wenn dem Medium 1% (w/v) Citrat zugesetzt wurde. Die Gene GOX0532 und GOX0536 kodieren für ExbB und einen Eisen- Siderophore Rezeptor, die beide wahrscheinlich an der Eisenaufnahme beteiligt sind. Die Supplementierung des Mediums mit 8 μM FeSO4 reduzierte die basale Expression beider Promotoren und erhöhte dadurch die Induktionsraten. Wahrscheinlich induziert durch Eisenlimitierung verursacht durch Citrate-Fe3+-Chelatierung, konnte die Expression von PGOX0532 und PGOX0536 >270-fach hochreguliert werden, wenn das Medium mit 5 x 50 mM Citrat in Intervallen von 1,5 Stunden versetzt wurden. ChAP-Seq Analysen bestätigten, dass der Transkriptionsfaktor Fur (ferric uptake regulator) (GOX0771) Promotorbereiche von PGOX0536 bindet.
Neben diesen Originalstudien wurde in einem Minireview ein Überblick über Expressionssysteme und ihre Anwendungen in Essigsäurebakterien veröffentlicht, der die Jahre von 1985, als die ersten Plasmide in Essigsäurebakterien verwendet wurden, bis Ende 2020 umfasste. In 6.097 Veröffentlichungen zu Essigsäurebakterien wurde lediglich für neun von 49 beschriebenen Gattungen über die Verwendung von Expressionsplasmiden berichtet. Innerhalb der beschriebenen Expressionsplasmide wurden sechs Hauptplasmidlinien identifiziert. Gemäß der Literaturrecherche sind die für G. oxydans entwickelten Systeme AraC-ParaBAD und TetR-Ptet die ersten regulierbaren Expressionssystem in dieser Bakteriengruppe, die eine >40-fach Induktion erlauben. Sie bieten neue Optionen für die zukünftige Stammentwicklung von G. oxydans und können möglicherweise auf andere Essigsäurebakterien-Arten übertragen werden.

Among industrially relevant microorganisms, the acetic acid bacterium Gluconobacter oxydans is valued for its ability to incompletely oxidize a vast number of carbohydrates stereo- and regio-specifically. Biotechnological production processes involving G. oxydans strains so far utilized exclusively constitutive target gene expression. Regulatable promoters used in G. oxydans suffered from low induction fold-changes and relatively high basal promoter activity already when not induced. This study aimed to establish regulatable promoter systems in G. oxydans that allow tuned target gene expression in an effector-dependent manner. For this purpose, expression plasmids were constructed to test four well-characterized heterologous regulator-promoter pairs in G. oxydans. Additionally, screenings were performed to identify regulatable endogenous G. oxydans promoters responding to a metabolite given as an external stimulus and suitable for controlled gene expression.
When transferred into G. oxydans, the AraC-ParaBAD system from Escherichia coli MC4100 permitted tight repression of target gene expression in the absence of L-arabinose and up to 480-fold induction when maximally induced with L-arabinose. At inducer concentrations from 0.1 to 1% (w/v) L-arabinose, reporter gene expression from ParaBAD was highly tunable. Furthermore, L-arabinose was found to be oxidized to L-arabinonic acid by the membrane-bound glucose dehydrogenase GdhM, resulting in an acidification of the medium and pH-dependent loss of intracellular reporter protein activity. This loss could be circumvented by pH-controlled cultivation or use of the gdhM deletion strain BP.6.
The oxidation product L-arabinonic acid is a valuable compound with potential use in various applications ranging from the production of pharmaceuticals to semi-conductor materials and composite cements. By using G. oxydans 621H, more than 120 g/L L-arabinonic acid could be produced from 160 g/L L-arabinose within 144 h in pH-controlled fed-batch cultivations. This high titer exceeded the ones previously reported for engineered E. coli or Saccharomyces cerevisiae strains more than threefold. Plasmid-based overexpression of gdhM to increase the enzyme activity did not improve the productivity possibly due to detrimental effects on the membrane integrity.
Besides the AraC-ParaBAD system, the suitability of the transposon Tn10-based TetR- Ptet system and of the LacI-PlacUV5 system, both originating from E. coli, was tested in G. oxydans. In contrast to AraC-ParaBAD, these inducible expression systems rely on target gene repression. The TetR-Ptet system outperformed the AraC-ParaBAD system, since in G. oxydans it showed extremely low basal expression and a concentration-dependent gradual increase of target gene expression upon induction with anhydrotetracycline. A maximal induction of up to 3,500-fold was obtained with 200 ng/mL anhydrotetracycline. Terminator sequences, plasmid backbone, antibiotic resistance gene, and the insertion of an additional known ribosome-binding site affected the performance of this system. In G. oxydans, only moderate inducible gene expression from PlacUV5 made this system the weakest of the regulator-promoter pairs tested in this work. Due to relatively high leakiness of non-induced PlacUV5, induction with 1 mM IPTG was only 40-fold.
With the RhaSR-PrhaBAD system from E. coli, a regulator-promoter pair was found that – contrary to E. coli – allows L-rhamnose-dependent knockdown of gene expression in G. oxydans instead of induction. PrhaBAD was very strong in G. oxydans and activated by the presence of RhaS in the absence of L-rhamnose, while with increasing L-rhamnose concentrations the PrhaBAD activity was reduced by 92% in a RhaS-dependent manner. In contrast, expression from the RhaS-dependent promoter PrhaT was L-rhamnose-inducible in G. oxydans as in E. coli. Since PrhaT was rather weak in G. oxydans, the induction was tunable up to 10-fold. This could be of advantage for expression of ‘difficult’ enzymes.
In the DNA microarray-based genome-wide stimulon screenings for regulated endogenous promotors, transcription of GOX0532 and GOX0536 encoding ExbB and a ferrisiderophore receptor both likely involved in iron uptake were found to be increased 45-fold when exposed to 1% (w/v) citrate as stimulus. Supplementation of the growth medium with 8 μM FeSO4 reduced the basal expression from both promoters und thereby increased induction ratios. Probably induced by iron-limitation through Fe3+-citrate chelation, expression from PGOX0532 and PGOX0536 could be upregulated >270-fold when 5 x 50 mM citrate was added to the medium in intervals of 1.5 h. ChAP-Seq experiments confirmed that the ferric uptake regulator Fur (GOX0771) binds to the promoter region PGOX0536.
Beside these original studies, a review on expression systems and their applications in acetic acid bacteria (AAB) was published covering the years from 1985, when the first plasmids were used in AAB, up to the end of 2020. Screening of 6097 AAB-related publications revealed that expression plasmids have been reported for merely nine out of 49 AAB genera currently described, and six major expression plasmid lineages were identified. According to this AAB-related literature search, the AraC-ParaBAD and TetR-Ptet systems developed by us for G. oxydans are the first regulatable expression system with induction ratios >40-fold in this group of bacteria. They provide new options in future strain development and may be transferable to other AAB species.
Quelle:Ahmer, B. M. M., Thomas, M. G., Larsen, R. A., & Postle, K. (1995). Characterization of the exbBD operon of Escherichia coli and the role of ExbB and ExbD in TonB function and stability. Journal of Bacteriology, 177(16), 4742–4747. doi:10.1128/jb.177.16.4742- 4747.1995
Ameyama, M., Matsushita, K., Shinagawa, E., & Adachi, O. (1987). Sugar-oxidizing respiratory chain of Gluconobacter suboxydans. Evidence for a branched respiratory chain and characterization of respiratory chain-linked cytochromes. Agricultural and Biological Chemistry, 51(11), 2943–2950. doi:10.1080/00021369.1987.10868527
Antoine, R., & Locht, C. (1992). Isolation and molecular characterization of a novel broad‐host‐ range plasmid from Bordetella bronchiseptica with sequence similarities to plasmids from Gram‐positive organisms. Molecular Microbiology, 6(13), 1785–1799. doi:10.1111/j.1365-2958.1992.tb01351.x
Aro-Kärkkäinen, N., Toivari, M., Maaheimo, H., & Ylilauri, M. (2014). L-Arabinose/D-galactose 1-dehydrogenase of Rhizobium leguminosarum bv. trifolii characterised and applied for bioconversion of L-arabinose to L-arabonate with Saccharomyces cerevisiae. Applied Micr, 9653–9665. doi:10.1007/s00253-014-6039-2
Bagg, A., & Neilands, J. B. (1987). Ferric uptake regulation protein acts as a repressor, employing iron(II) as a cofactor to bind the operator of an iron transport operon in Escherichia coli, (1970), 5471–5477. doi:10.1021/bi00391a039
Bailey, T. L., & Gribskov, M. (1998). Combining evidence using p-values: Application to sequence homology searches. Bioinformatics, 14(1), 48–54. doi:10.1093/bioinformatics/14.1.48
Barton, H. A., Johnson, Z., Cox, C. D., Vasil, A. I., & Vasil, M. L. (1996). Ferric uptake regulator mutants of Pseudomonas aeruginosa with distinct alterations in the iron-dependent repression of exotoxin A and siderophores in aerobic and microaerobic environments. Molecular Microbiology, 21(5), 1001–1017. doi:10.1046/j.1365-2958.1996.381426.x
Battling, S., Wohlers, K., Igwe, C., Kranz, A., Pesch, M., Wirtz, A., ... Bott, M. (2020). Novel plasmid-free Gluconobacter oxydans strains for production of the natural sweetener 5- ketofructose. Microbial Cell Factories, 19(1), 1–10. doi:10.1186/s12934-020-01310-7
Bertram, R., & Hillen, W. (2008). The application of Tet repressor in prokaryotic gene regulation and expression. Microbial Biotechnology, 1, 2–16. doi:10.1111/j.1751- 7915.2007.00001.x
Bhende, P. M., & Egan, S. M. (2000). Genetic evidence that transcription activation by RhaS involves specific amino acid contacts with sigma 70. Journal of Bacteriology, 182(17), 4959–4969. doi:10.1128/JB.182.17.4959-4969.2000
Blank, M., & Schweiger, P. (2018). Surface display for metabolic engineering of industrially important acetic acid bacteria. PeerJ, 6, 1–19. doi:10.7717/peerj.4626
Braun, V. (2003). Iron uptake by Escherichia coli. Frontiers in Bioscience, 1409–1421. doi:10.2741/1232
Brautaset, T., Lale, R., & Valla, S. (2009). Positively regulated bacterial expression systems. Microbial Biotechnology, 2(1), 15–30. doi:10.1111/j.1751-7915.2008.00048.x
Bringer, S., & Bott, M. (2016). Central carbon Mmetabolism and respiration in Gluconobacter oxydans. In Acetic Acid Bacteria (pp. 235–253). doi:10.1007/978-4-431-55933-7_11
Brown, N. L., Stoyanov, J. V., Kidd, S. P., & Hobman, J. L. (2003). The MerR family of transcriptional regulators. FEMS Microbiology Reviews, 27(2–3), 145–163. doi:10.1016/S0168-6445(03)00051-2
Browning, D. F., & Busby, S. J. W. (2004). The regulation of bacterial transcription initiation. Nature Reviews Microbiology, 2(1), 57–65. doi:10.1038/nrmicro787
Browning, D. F., & Busby, S. J. W. (2016). Local and global regulation of transcription initiation in bacteria. Nature Reviews Microbiology, 14(10), 638–650. doi:10.1038/nrmicro.2016.103
Bugala, J., Cimova, V., Grones, P., & Grones, J. (2016). Characterization of newly identified DnaA and DnaB proteins from Acetobacter. Research in Microbiology, 167(8), 655–668. doi:10.1016/j.resmic.2016.06.010
Celia, H., Noinaj, N., & Buchanan, S. K. (2020). Structure and stoichiometry of the Ton molecular motor. International Journal of Molecular Sciences, 21(2), 1–15. doi:10.3390/ijms21020375
Chen, R. R. (2007). Permeability issues in whole-cell bioprocesses and cellular membrane engineering. Applied Microbiology and Biotechnology, 74(4), 730–738. doi:10.1007/s00253-006-0811-x
Condon, C., FitzGerald, R. J., & O’Gara, F. (1991). Conjugation and heterologous gene expression in Gluconobacter oxydans ssp. suboxydans. FEMS Microbiology Letters, 80, 173–178. doi:10.1111/j.1574-6968.1991.tb04656.x
Dauner, M., Sonderegger, M., Hochuli, M., Szyperski, T., Wüthrich, K., Hohmann, H. P., ... Bailey, J. E. (2002). Intracellular carbon fluxes in riboflavin-producing Bacillus subtilis during growth on two-carbon substrate mixtures. Applied and Environmental Microbiology, 68(4), 1760–1771. doi:10.1128/AEM.68.4.1760-1771.2002
De Muynck, C., Pereira, C. S. S., Naessens, M., Parmentier, S., Soetaert, W., & Vandamme, E. J. (2007). The genus Gluconobacter oxydans: Comprehensive overview of biochemistry and biotechnological applications. Critical Reviews in Biotechnology, 27(3), 147–171. doi:10.1080/07388550701503584
Degenkolb, J., Takahashi, M., Ellestad, G. a, & Hillen, W. (1991). Structural requirements of tetracycline-Tet repressor interaction: Determination of equilibrium binding constants for tetracycline analogs with the Tet repressor. Antimicrobial Agents and Chemotherapy, 35(8), 1591. doi:10.1128/aac.35.8.1591
Deppenmeier, U., Hoffmeister, M., & Prust, C. (2002). Biochemistry and biotechnological applications of Gluconobacter strains. Applied Microbiology and Biotechnology, 60(3), 233–242. doi:10.1007/s00253-002-1114-5.
Dunn, T. M., & Schleif, R. (1984). Deletion analysis of the Escherichia coli ara PC and PBAD promoters. Journal of Molecular Biology, 180(1), 201–204. doi:10.1016/0022- 2836(84)90437-6
Ebright, R. H., & Busby, S. (1995). The Escherichia coli RNA polymerase α-subunit: structure and function. Current Opinion in Genetics and Development, 5(2), 197–203. doi:10.1016/0959-437X(95)80008-5
Egan, S. M., & Schleif, R. F. (1993). A regulatory cascade in the induction of rhaBAD. Journal of Molecular Biology. doi:10.1006/jmbi.1993.1565
Ernst, F. D., Bereswill, S., Waidner, B., Stoof, J., Mäder, U., Kusters, J. G., ... Homuth, G. (2005). Transcriptional profiling of Helicobacter pylori Fur- and iron-regulated gene expression. Microbiology, 151(2), 533–546. doi:10.1099/mic.0.27404-0
Escolar, L., Pérez-Martín, J., & De Lorenzo, V. (1998). Binding of the Fur (ferric uptake regulator) repressor of Escherichia coli to arrays of the GATAAT sequence. Journal of Molecular Biology, 283(3), 537–547. doi:10.1006/jmbi.1998.2119
Fernández-Castané, A., Vine, C. E., Caminal, G., & López-Santín, J. (2012). Evidencing the role of lactose permease in IPTG uptake by Escherichia coli in fed-batch high cell density cultures. Journal of Biotechnology, 157(3), 391–398. doi:10.1016/j.jbiotec.2011.12.007
Fillat, M. F. (2014). The FUR (ferric uptake regulator) superfamily: Diversity and versatility of key transcriptional regulators. Archives of Biochemistry and Biophysics, 546, 41–52. doi:10.1016/j.abb.2014.01.029
Florea, M., Santosa, G., Hagemann, H., Reeve, B., Abbott, J., Freemont, P. S., ... Spencer- Milnes, X. (2016). Engineering control of bacterial cellulose production using a genetic toolkit and a new cellulose-producing strain. Proceedings of the National Academy of Sciences, 113(24), E3431–E3440. doi:10.1073/pnas.1522985113
Fricke, P. M., Klemm, A., Bott, M., & Polen, T. (2021a). On the way toward regulatable expression systems in acetic acid bacteria: Target gene expression and use cases. Applied Microbiology and Biotechnology. doi:10. 1007/ s00253- 021- 11269-z
Fricke, P. M., Link, T., Gätgens, J., Sonntag, C., Otto, M., Bott, M., & Polen, T. (2020). A tunable L-arabinose-inducible expression plasmid for the acetic acid bacterium Gluconobacter oxydans. Applied Microbiology and Biotechnology. doi:10.1007/s00253- 020-10905-4
Fricke, P. M., Lürkens, M., Hünnefeld, M., Sonntag, C. K., Bott, M., Davari, M. D., & Polen, T. (2021b). Highly tunable TetR‐dependent target gene expression in the acetic acid bacterium Gluconobacter oxydans. Applied Microbiology and Biotechnology. doi:10.1007/s00253-021-11473-x
Fukaya, M., Okumura, H., Masai, H., Uozumi, T., & Beppu, T. (1985). Development of a host- vector system for Gluconobacter suboxydans. Agricultural and Biological Chemistry, 49(8), 2407–2411. doi:10.1080/00021369.1985.10867095
Gätgens, C., Degner, U., Bringer-Meyer, S., & Herrmann, U. (2007). Biotransformation of glycerol to dihydroxyacetone by recombinant Gluconobacter oxydans DSM 2343. Applied Microbiology and Biotechnology, 76(3), 553–559. doi:10.1007/s00253-007- 1003-z
Greenblatt, J., Li, J., Adhyat, S., Friedmanf, D. I., Baron, L. S., Redfieldt, B., ... Weissbachi, H. (1980). L factor that is required for β-galactosidase synthesis is the nusA gene product involved in transcription termination. Proc.Natl.Acad.Sci.U.S.A, 77(4), 1991–1994. doi:10.1073/pnas.77.4.1991
Grkovic, S., Brown, M. H., & Skurray, R. A. (2002). Regulation of bacterial drug export systems. Microbiology and Molecular Biology Reviews, 66(4), 671–701. doi:10.1128/MMBR.66.4.671-701.2002
Gross, J., & Englesberg, E. (1959). Determination of the order of mutational sites governing L- arabinose utilization in Escherichia coli B/r by transduction with phage P1bt. Virology, 9(3), 314–331. doi:10.1016/0042-6822(59)90125-4
Gupta, A., Singh, V. K., Qazi, G. N., & Kumar, A. (2001). Gluconobacter oxydans: Its biotechnological applications. Journal of Molecular Microbiology and Biotechnology, 3(3), 445–456.
Guzman, L., Belin, D., Carson, M. J., & Beckwith, J. (1995). Tight regulation, modulation, and high-level expression by vectors containing the arabinose PBAD promoter. Journal of Applied Microbiology, 177(14), 4121–4130. doi:10.1128/jb.177.14.4121-4130.1995
Hansen, L. H., Knudsen, S., & Sørensen, S. J. (1998). The effect of the lacY gene on the induction of IPTG inducible promoters, studied in Escherichia coli and Pseudomonas fluorescens. Current Microbiology, 36(6), 341–347. doi:10.1007/s002849900320
Hahn, S., & Schleif, R. (1983). In vivo regulation of the Escherichia coli araC promoter. Journal of Bacteriology, 155(2), 593–600. doi:10.1128/jb.155.2.593-600.1983
Hahn, S., Dunn, T., & Schleif, R. (1984). Upstream repression and CRP stimulation of the Escherichia coli L-arabinose operon. Journal of Molecular Biology, 180(1), 61–72. doi:10.1016/0022-2836(84)90430-3
Hakimi, M. A., Privat, I., Valay, J. G., & Lerbs-Mache, S. (2000). Evolutionary conservation of C-terminal domains of primary sigma70-type transcription factors between plants and bacteria. Journal of Biological Chemistry, 275(13), 9215–9221. doi:10.1074/jbc.275.13.9215
Hanke, T., Nöh, K., Noack, S., Polen, T., Bringer, S., Sahm, H., ... Bott, M. (2013). Combined fluxomics and transcriptomics analysis of glucose catabolism via a partially cyclic pentose phosphate pathway in Gluconobacter oxydans 621H. Applied and Environmental Microbiology, 79(7), 2336–2348. doi:10.1128/AEM.03414-12
Hanke, T., Richhardt, J., Polen, T., Sahm, H., Bringer, S., & Bott, M. (2012). Influence of oxygen limitation, absence of the cytochrome bc1 complex and low pH on global gene expression in Gluconobacter oxydans 621H using DNA microarray technology. Journal of Biotechnology, 157(3), 359–372. doi:10.1016/j.jbiotec.2011.12.020
Hantke, K. (1981). Regulation of ferric iron transport in Escherichia coli K12: Isolation of a constitutive mutant. MGG Molecular & General Genetics, 182(2), 288–292. doi:10.1007/BF00269672
Harmer, T., Wu, M., & Schleif, R. (2001). The role of rigidity in DNA looping-unlooping by AraC. Proceedings of the National Academy of Sciences of the United States of America, 98(2), 427–431. doi:10.1073/pnas.98.2.427
Haugen, S. P., Ross, W., & Gourse, R. L. (2008). Advances in bacterial promoter recognition and its control by factors that do not bind DNA. Nature Reviews Microbiology, 6(7), 507– 519. doi:10.1038/nrmicro1912
Herweg, E., Schöpping, M., Rohr, K., Siemen, A., Frank, O., Hofmann, T., ... Büchs, J. (2018). Production of the potential sweetener 5-ketofructose from fructose in fed-batch cultivation with Gluconobacter oxydans. Bioresource Technology, 259(March), 164–172. doi:10.1016/j.biortech.2018.03.038
Herzenberg, L. A. (1961). Isolotion and identification of derivatives formed in the course of intracellular accumulation of thiogalactosides by Escherichia coli. Archives of Biochemistry and Biophysics, 93(2), 314–315. doi:10.1016/0003-9861(61)90270-3
Higgs, P. I., Larsen, R. A., & Postle, K. (2002). Quantification of known components of the Escherichia coli TonB energy transduction system: TonB, ExbB, ExbD and FepA. Molecular Microbiology, 44(1), 271–281. doi:10.1046/j.1365-2958.2002.02880.x
Hirschel, B. J., Shen, V., & Schlessinger, D. (1980). Lactose operon transcription from wild- type and L8-UV5 lac promoters in Escherichia coli treated with chloramphenicol. Journal of Bacteriology, 143(3), 1534–1537. doi:10.1128/jb.143.3.1534-1537.1980
Hoffmann, J. J., Hövels, M., Kosciow, K., & Deppenmeier, U. (2020). Synthesis of the alternative sweetener 5-ketofructose from sucrose by fructose dehydrogenase and invertase producing Gluconobacter strains. Journal of Biotechnology, 307(October 2019), 164–174. doi:10.1016/j.jbiotec.2019.11.001
Holcroft, C. C., & Egan, S. M. (2000). Interdependence of activation at rhaSR by cyclic AMP receptor protein, the RNA polymerase alpha subunit C-terminal domain, and RhaR. Journal of Bacteriology, 182(23), 6774–6782. doi:10.1128/JB.182.23.6774-6782.2000
Horazdovsky, B. F., & Hogg, R. W. (1989). Genetic reconstitution of the high-affinity L- arabinose transport system. Journal of Bacteriology, 171(6), 3053–3059. doi:10.1128/jb.171.6.3053-3059.1989
Horowitz, H., & Platt, T. (1982). A termination site for lacI transcription is between the CAP site and the lac promoter. Journal of Biological Chemistry, 257(19), 11740–11746. doi:10.1016/s0021-9258(18)33825-0
Hu, Y., Wan, H., Li, J., & Zhou, J. (2015). Enhanced production of L-sorbose in an industrial Gluconobacter oxydans strain by identification of a strong promoter based on proteomics analysis. Journal of Industrial Microbiology and Biotechnology, 42(7), 1039–1047. doi:10.1007/s10295-015-1624-7
Jacob, F., & Monod, J. (1961). Genetic regulatory mechanisms in the synthesis of proteins. Journal of Molecular Biology, 3(3), 318–356. doi:10.1016/S0022-2836(61)80072-7
Jensen, P. R., Westerhoff, H. V., & Michelsen, O. (1993). The use of lac‐type promoters in control analysis. European Journal of Biochemistry, 211(1–2), 181–191. doi:10.1111/j.1432-1033.1993.tb19885.x
Kallnik, V. (2012). Untersuchungen zur Polyoloxidation in Gluconobacter oxydans und Thermotoga maritima.
Kallnik, V., Meyer, M., Deppenmeier, U., & Schweiger, P. (2010). Construction of expression vectors for protein production in Gluconobacter oxydans. Journal of Biotechnology, 150(4), 460–465. doi:10.1016/j.jbiotec.2010.10.069
Kania, J., & Brown, D. T. (1976). The functional repressor parts of a tetrameric lac repressor- β-galactosidase chimaera are organized as dimers. Proceedings of the National Academy of Sciences of the United States of America, 73(10), 3529–3533. doi:10.1073/pnas.73.10.3529
Kelly, C. L., Taylor, G. M., Hitchcock, A., Torres-Méndez, A., & Heap, J. T. (2018). A rhamnose- inducible system for precise and temporal control of gene expression in Cyanobacteria. ACS Synthetic Biology, 7(4), 1056–1066. doi:10.1021/acssynbio.7b00435
Kersters, K., & De Ley, J. (1968). The occurrence of the Entner-Doudoroff pathway in bacteria. Antonie van Leeuwenhoek, 34(1), 393–408. doi:10.1007/BF02046462
Kersters, K., Linsdiyanti, P., Komagata, K., & Swings, J. (2006). The family Acetobacteraceae: The genera Acetobacter, Acidomonas, Asaia, Gluconacetobacter, Gluconobacter, and Kozakia. Prokaryotes.
Kiefler, I. (2016). Strain development of Gluconobacter oxydans: Complementation of non- functional metabolic pathways and increase of carbon flux.
Kiefler, I., Bringer, S., & Bott, M. (2015). SdhE-dependent formation of a functional Acetobacter pasteurianus succinate dehydrogenase in Gluconobacter oxydans − a first step toward a complete tricarboxylic acid cycle. Applied Microbiology and Biotechnology, 99(21), 9147–9160. doi:10.1007/s00253-015-6972-8
Kiefler, I., Bringer, S., & Bott, M. (2017). Metabolic engineering of Gluconobacter oxydans 621H for increased biomass yield. Applied Microbiology and Biotechnology, 5453–5467. doi:10.1007/s00253-017-8308-3
Kingston, R. E., & Chamberlin, M. J. (1981). Pausing and attenuation of in vitro transcription in the rrnB operon of E. coli. Cell, 27(3 PART 2), 523–531. doi:10.1016/0092- 8674(81)90394-9
Kitova, A., Tarasov, S., Plekhanova, Y., Bykov, A., & Reshetilov, A. (2021). Direct bioelectrocatalytic oxidation of glucose by Gluconobacter oxydans membrane fractions in PEDOT:PSS/TEG-modified biosensors. Biosensors, 11(5), 144. doi:10.3390/bios11050144
Kleinschmidt, C., Tovar, K., & Hillen, W. (1991). Computer simulations and experimental studies of gel mobility patterns for weak and strong non-cooperative protein binding to two targets on the same DNA: Application to binding of Tet repressor variants to multiple and single tet operator sites. Nucleic Acids Research, 19(5), 1021–1028. doi:10.1093/nar/19.5.1021
Kolin, A., Balasubramaniam, V., Skredenske, J. M., Wickstrum, J. R., & Egan, S. M. (2008). Differences in the mechanism of the allosteric L-rhamnose responses of the AraC/XylS family transcription activators RhaS and RhaR. Molecular Microbiology, 68(2), 448–461. doi:10.1111/j.1365-2958.2008.06164.x
Kolodrubetz, D., & Schleif, R. (1981). Regulation of the L-arabinose transport operons in Escherichia coli. Journal of Molecular Biology, 151(2), 215–227. doi:10.1016/0022- 2836(81)90512-X
Kondo, K., & Horinouchi, S. (1997). Characterization of the genes encoding the three- component membrane-bound alcohol dehydrogenase from Gluconobacter suboxydans and their expression in Acetobacter pasteurianus. Applied and Environmental Microbiology, 63(3), 1131–1138. doi:10.1128/aem.63.3.1131-1138.1997
Kosciow, K., Zahid, N., Schweiger, P., & Deppenmeier, U. (2014). Production of a periplasmic trehalase in Gluconobacter oxydans and growth on trehalose. Journal of Biotechnology, 189, 27–35. doi:10.1016/j.jbiotec.2014.08.029
Kosciow, K., Domin, C., Schweiger, P., & Deppenmeier, U. (2016). Extracellular targeting of an active endoxylanase by a TolB negative mutant of Gluconobacter oxydans. Journal of Industrial Microbiology & Biotechnology, 43(7), 989–999. doi:10.1007/s10295-016- 1770-6
Kostner, D., Peters, B., Mientus, M., Liebl, W., & Ehrenreich, A. (2013). Importance of codB for new codA-based markerless gene deletion in Gluconobacter strains. Applied Microbiology and Biotechnology, 97(18), 8341–8349. doi:10.1007/s00253-013-5164-7
Kovach, M. E., Elzer, P. H., Hill, D. S., Robertson, G. T., Farris, M. A., Roop, R. M., & Peterson, K. M. (1995). Four new derivatives of the broad-host-range cloning vector pBBR1MCS, carrying different antibiotic resistance cassettes. Gene, 166(1), 175–176. doi:10.1089/152045500436104
Krajewski, V., Simic, P., Mouncey, N. J., Bringer, S., Sahm, H., & Bott, M. (2010). Metabolic engineering of Gluconobacter oxydans for improved growth rate and growth yield on glucose by elimination of gluconate formation, 76(13), 4369–4376. doi:10.1128/AEM.03022-09
Kranz, A., Vogel, A., Degner, U., Kiefler, I., Bott, M., Usadel, B., & Polen, T. (2017). High precision genome sequencing of engineered Gluconobacter oxydans 621H by combining long nanopore and short accurate Illumina reads. Journal of Biotechnology, 258(January), 197–205. doi:10.1016/j.jbiotec.2017.04.016
Kulhánek, M. (1989). Microbial dehydrogenations of monosaccharides. Advances in Applied Microbiology, 34, 141–182. doi:10.1016/S0065-2164(08)70318-6
Lavrrar, J. L., & McIntosh, M. A. (2003). Architecture of a Fur binding site: A comparative analysis. Journal of Bacteriology, 185(7), 2194–2202. doi:10.1128/JB.185.7.2194- 2202.2003
Lee, D. J., Minchin, S. D., & Busby, S. J. W. (2012). Activating transcription in bacteria. Annual Review of Microbiology, 66, 125–152. doi:10.1146/annurev-micro-092611-150012
Ley, J. De, Gillis, M., & Swings, J. (1984). The genus Gluconobacter. In N. R. Krieg & J. G. Holt (Eds.), Bergey’s Manual of Systematic Bacteriology (pp. 267–278).
Liu, H., Valdehuesa, K. N. G., Ramos, K. R. M., Nisola, G. M., Lee, W.-K., & Chung, W.-J. (2014). L-arabinoate and D-galactonate production by expressing a versatile sugar dehydrogenase in metabolically engineered Escherichia coli. Bioresource Technology, 159, 455–459. doi:10.1016/j.biortech.2014.03.056.
Liu, L. P., Yang, X., Zhao, X. J., Zhang, K. Y., Li, W. C., Xie, Y. Y., ... Zhong, C. (2020). A lambda red and FLP/FRT-mediated site-specific recombination system in Komagataeibacter xylinus and its application to enhance the productivity of bacterial cellulose. ACS Synthetic Biology, 9(11), 3171–3180. doi:10.1021/acssynbio.0c00450
Lobell, R. B., & Schleif, R. F. (1991). AraC-DNA looping: Orientation and distance-dependent loop breaking by the cyclic AMP receptor protein. Journal of Molecular Biology, 218(1), 45–54. doi:10.1016/0022-2836(91)90872-4
Lonetto, M. A., Rhodius, V., Lamberg, K., Kiley, P., Busby, S., & Gross, C. (1998). Identification of a contact site for different transcription activators in region 4 of the Escherichia coli RNA polymerase σ70 subunit. Journal of Molecular Biology, 284(5), 1353–1365. doi:10.1006/jmbi.1998.2268
Marbach, A., & Bettenbrock, K. (2012). lac operon induction in Escherichia coli: Systematic comparison of IPTG and TMG induction and influence of the transacetylase LacA. Journal of Biotechnology, 157(1), 82–88. doi:10.1016/j.jbiotec.2011.10.009
Matsushita, K., Toyama, H., & Adachi, O. (1994). Respiratory chains and bioenergetics of acetic acid bacteria. Advances in Microbial Physiology, 36, 247–301. doi:10.1016/s0065- 2911(08)60181-2.
McMurry, L., Petrucci, R. E., & Levy, S. B. (1980). Active efflux of tetracycline encoded by four genetically different tetracyline resistance determinants in Escherichia coli. Proceedings of the National Academy of Sciences of the United States of America, 77(7 II), 3974– 3977. doi:10.1073/pnas.77.7.3974
Meier, I., Wray, L. V., & Hillen, W. (1988). Differential regulation of the Tn10-encoded tetracycline resistance genes tetA and tetR by the tandem tet operators O1 and O2. The EMBO Journal, 7(2), 567–572. doi:10.1002/j.1460-2075.1988.tb02846.x
Merfort, M., Herrmann, U., Bringer-Meyer, S., & Sahm, H. (2006). High-yield 5-keto-D-gluconic acid formation is mediated by soluble and membrane-bound gluconate-5- dehydrogenases of Gluconobacter oxydans. Applied Microbiology and Biotechnology, 73, 443–451. doi:10.1007/s00253-006-0467-6
Merfort, Marcel, Herrmann, U., Ha, S. W., Elfari, M., Bringer-Meyer, S., Görisch, H., & Sahm, H. (2006). Modification of the membrane-bound glucose oxidation system in Gluconobacter oxydans significantly increases gluconate and 5-keto-D-gluconic acid accumulation. Biotechnology Journal, 1(5), 556–563. doi:10.1002/biot.200600032
Mergulhão, F. J. M., & Monteiro, G. A. (2004). Secretion capacity limitations of the Sec pathway in Escherichia coli. Journal of Microbiology and Biotechnology, 128–133.
Meyer, M., Schweiger, P., & Deppenmeier, U. (2013). Effects of membrane-bound glucose dehydrogenase overproduction on the respiratory chain of Gluconobacter oxydans. Applied Microbiology and Biotechnology, 97(8), 3457–3466. doi:10.1007/s00253-012- 4265-z
Mientus, M., Kostner, D., Peters, B., Liebl, W., & Ehrenreich, A. (2017). Characterization of membrane-bound dehydrogenases of Gluconobacter oxydans 621H using a new system for their functional expression. Applied Microbiology and Biotechnology, 3189–3200. doi:10.1007/s00253-016-8069-4
Mill, J. H. (1992). A short course in bacterial genetics. Cold Spring Harbor: Cold Spring Harbor Laboratory Press.
Müller-Hill, B. (1998). The function of auxiliary operators. Molecular Microbiology, 29(1), 13– 18. doi:10.1046/j.1365-2958.1998.00870.x
Nguyen, T. N. M., Phan, Q. G., Duong, L. P., Bertrand, K. P., & Lenski, R. E. (1989). Effects of carriage and expression of the Tn10 tetracycline-resistance operon on the fitness of Escherichia coli K12. Molecular Biology and Evolution, 6(3), 213–225. doi:10.1093/oxfordjournals.molbev.a040545
Noel, R. J., & Reznikoff, W. S. (2000). Structural studies of lacUV5-RNA polymerase interactions in vitro. Ethylation interference and missing nucleoside analysis. Journal of Biological Chemistry, 275(11), 7708–7712. doi:10.1074/jbc.275.11.7708
Oehler, S., Eismann, E. R., Kramer, H., & Muller-Hill, B. (1990). The three operators of the lac operon cooperate in repression. EMBO Journal, 9(4), 973–979. doi:10.1002/j.1460- 2075.1990.tb08199.x
Ohrem, H. L., & Voß, H. (1996). Inhibitory effects of glycerol on Gluconobacter oxydans. Biotechnology Letters, 18(3), 245–250. doi:10.1007/bf00142939
Olijve, W., & Kok, J. J. (1979). An analysis of the growth of Gluconobacter oxydans in chemostat cultures. Archives of Microbiology, 121(3), 291–297. doi:10.1007/BF00425070
Oliva, B., Gordon, G., Mcnicholas, P., Ellestad, G., & Choprat, I. a N. (1992). Evidence that tetracycline analogs whose primary target is not the bacterial ribosome cause lysis of Escherichia coli. Antimicrobial Agents and Chemotherapy, 36(5), 913–919. doi:10.1128/AAC.36.5.913
Paget, M., & Helmann, J. (2003). The σ70 family of sigma factors. Genome Biology, 4(1), 1–6. doi:10.1186/gb-2003-4-1-203.
Paget, M. S. (2015). Bacterial sigma factors and anti-sigma factors: Structure, function and distribution. Biomolecules, 5(3), 1245–1265. doi:10.3390/biom5031245
Pappenberger, G., & Hohmann, H.-P. (2014). Industrial production of L-ascorbic acid (vitamin C) and D-isoascorbic acid. Advances in Biochemical Engineering/Biotechnology, 143, 143–188. doi:10.1007/10_2013_243
Pasqua, M., Visaggio, D., Sciuto, A. Lo, Genah, S., Banin, E., Visca, P., & Imperi, F. (2017). Ferric uptake regulator Fur is conditionally essential in Pseudomonas aeruginosa. Journal of Bacteriology, 199(22), 1–17. doi:10.1128/JB.00472-17
Peters, B., Mientus, M., Kostner, D., Junker, A., Liebl, W., Ehrenreich, A., & Received: (2013). Characterization of membrane-bound dehydrogenases from Gluconobacter oxydans 621H via whole-cell activity assays using multideletion strains. Applied Microbiology and Biotechnology, 97, 6397–6412. doi:10.1007/s00253-013-4824-y
Philips, S. J., Canalizo-Hernandez, M., Yildirim, I., Schatz, G. C., Mondragón, A., & O’Halloran, T. V. (2015). Allosteric transcriptional regulation via changes in the overall topology of the core promoter. Science, 349(6250), 877–881. doi:10.1126/science.aaa9809
Platt, T. (1981). Termination of transcription and its regulation in the tryptophan operon of E. coli. Cell, 24(1), 10–23. doi:10.1016/0092-8674(81)90496-7
Possoz, C., Newmark, J., Sorto, N., Sherratt, D. J., & Tolmasky, M. E. (2007). Sublethal concentrations of the aminoglycoside amikacin interfere with cell division without affecting chromosome dynamics. Antimicrobial Agents and Chemotherapy, 51(1), 252– 256. doi:10.1128/AAC.00892-06
Power, J. (1967). The L-rhamnose genetic system in Escherichia coli K-12. Genetics, 55(3), 557–568. doi:10.1093/genetics/55.3.557
Pramanik, A., Zhang, F., Schwarz, H., Schreiber, F., & Braun, V. (2010). ExbB protein in the cytoplasmic membrane of Escherichia coli forms a stable oligomer. Biochemistry, 8721– 8728. doi:10.1021/bi101143y
Prust, C., Hoffmeister, M., Liesegang, H., Wiezer, A., Fricke, W. F., Ehrenreich, A., ... Deppenmeier, U. (2005). Complete genome sequence of the acetic acid bacterium Gluconobacter oxydans. Nature Biotechnology, 23(2), 195–200. doi:10.1038/nbt1062
Raspor, P., & Goranovič, D. (2008). Biotechnological applications of acetic acid bacteria. Critical Reviews in Biotechnology, 28(2), 101–124. doi:10.1080/07388550802046749
Reznikoff, W. S. (1992). The lactose operon‐controlling elements: A complex paradigm. Molecular Microbiology, 6(17), 2419–2422. doi:10.1111/j.1365-2958.1992.tb01416.x
Richhardt, J., Bringer, S., & Bott, M. (2012). Mutational analysis of the pentose phosphate and Entner-Doudoroff pathways in Gluconobacter oxydans reveals improved growth of a Δedd Δeda mutant on mannitol. Applied and Environmental Microbiology, 78(19), 6975– 6986. doi:10.1128/AEM.01166-12
Richhardt, J., Bringer, S., & Bott, M. (2013). Role of the pentose phosphate pathway and the Entner-Doudoroff pathway in glucose metabolism of Gluconobacter oxydans 621H. Applied Microbiology Biotechnology, 97(10), 4315–4323. doi:10.1007/s00253-013- 4707-2
Richhardt, J., Luchterhand, B., Bringer, S., Büchs, J., & Bott, M. (2013). Evidence for a key role of cytochrome bo3 oxidase in respiratory energy metabolism of Gluconobacter oxydans. Journal of Bacteriology, 195(18), 4210–4220. doi:10.1128/JB.00470-13
Rosenberg, H. F. (1998). Isolation of recombinant secretory proteins by limited induction and quantitative harvest. BioTechniques, 24(2), 188–192. doi:10.2144/98242bm03
Saïda, F., Uzan, M. ., Odaert, B., & Bontems, F. (2006). Expression of highly toxic genes in E. coli: Special strategies and genetic tools. Current Protein and Peptide Science, 7, 47– 56. doi:10.2174/138920306775474095
Saito, Y., Ishii, Y., Hayashi, H., Imao, Y., Akashi, T., Yoshikawa, K., ... Shimomura, K. (1997). Cloning of genes coding for L-sorbose and L-sorbosone dehydrogenases from Gluconobacter oxydans and microbial production of 2-keto-L-gulonate, a precursor of L- ascorbic acid, in a recombinant G. oxydans strain. Applied and Environmental Microbiology, 63(2), 454–460. doi:10.1128/aem.63.2.454-460.1997
Santillán, M., & Mackey, M. C. (2008). Quantitative approaches to the study of bistability in the lac operon of Escherichia coli. Journal of the Royal Society Interface, 5, S29–S39. doi:10.1098/rsif.2008.0086.focus
Sarvan, S., Butcher, J., Stintzi, A., & Couture, J. F. (2018). Variation on a theme: Investigating the structural repertoires used by ferric uptake regulators to control gene expression. BioMetals, 31(5), 681–704. doi:10.1007/s10534-018-0120-8
Sarvan, S., Yeung, A., Charih, F., Stintzi, A., & Couture, J. F. (2019). Purification and characterization of Campylobacter jejuni ferric uptake regulator. BioMetals, 32(3), 491– 500. doi:10.1007/s10534-019-00177-5
Schauer, K., Rodionov, D. A., & de Reuse, H. (2008). New substrates for TonB-dependent transport: Do we only see the “tip of the iceberg”? Trends in Biochemical Sciences, 33(7), 330–338. doi:10.1016/j.tibs.2008.04.012
Schedel, M. (2001). Regioselective oxidation of aminosorbitol with Gluconobacter oxydans, key reaction in the industrial 1-deoxynojirimycin synthesis (Vol. 8), 295–311.
Schleif, R. (2010). AraC protein, regulation of the L-arabinose operon in Escherichia coli, and the light switch mechanism of AraC action. FEMS Microbiology Reviews, 34(5), 779– 796. doi:10.1111/j.1574-6976.2010.00226.x
Schweikert, S., Kranz, A., Yakushi, T., Filipchyk, A., Polen, T., Etterich, H., ... Bott, M. (2021). The FNR-type regulator GoxR of the obligatory aerobic acetic acid bacterium Gluconobacter oxydans affects expression of genes involved in respiration and redox metabolism. Applied and Environmental Microbiology, (January), 1–9. doi:10.1128/aem.00195-21
Semsey, S., Tolstorukov, M. Y., Virnik, K., Zhurkin, V. B., & Adhya, S. (2004). DNA trajectory in the Gal repressosome. Genes and Development, 18(15), 1898–1907. doi:10.1101/gad.1209404
Shi, L., Li, K., Zhang, H., Liu, X., Lin, J., & Wei, D. (2014). Identification of a novel promoter gHp0169 for gene expression in Gluconobacter oxydans. Journal of Biotechnology, 175, 69–74. doi:10.1016/j.jbiotec.2014.01.035
Siemen, A., Kosciow, K., Schweiger, P., & Deppenmeier, U. (2018). Production of 5- ketofructose from fructose or sucrose using genetically modified Gluconobacter oxydans strains. Applied Microbiology and Biotechnology, 102(4), 1699–1710. doi:10.1007/s00253-017-8699-1
Silverstone, A. E., Arditti, R. R., & Magasanik, B. (1970). Catabolite-insensitive revertants of lac promoter mutants. Proceedings of the National Academy of Sciences of the United States of America, 66(3), 773–779. doi:10.1073/pnas.66.3.773
Soini, J., Ukkonen, K., & Neubauer, P. (2008). High cell density media for Escherichia coli are generally designed for aerobic cultivations - consequences for large-scale bioprocesses and shake flask cultures. Microbial Cell Factories, 7, 1–11. doi:10.1186/1475-2859-7-26
Soisson, S. M., MacDougall-Shackleton, B., Schleif, R., & Wolberger, C. (1997). Structural basis for ligand-regulated oligomerization of AraC. Science, 276(5311), 421–425. doi:10.1126/science.276.5311.421
Studier, F. W., & Moffatt, B. A. (1986). Use of bacteriophage T7 RNA polymerase to direct selective high-level expression of cloned genes. Journal of Molecular Biology, 189(1), 113–130. doi:10.1016/0022-2836(86)90385-2
Švitel, J., Tkáč, J., Voštiar, I., Navrátil, M., Štefuca, V., Bučko, M., & Gemeiner, P. (2006). Gluconobacter in biosensors: Applications of whole cells and enzymes isolated from Gluconobacter and Acetobacter to biosensor construction. Biotechnology Letters, 28(24), 2003–2010. doi:10.1007/s10529-006-9195-3
Swint-Kruse, L. and Matthews, K. S. (2009). Allostery in the LacI/GalR family: Variations on a theme. Current Opinion Microbiology, 12(2), 129–137. doi:10.1016/j.mib.2009.01.009
Tang, S. Y., Fazelinia, H., & Cirino, P. C. (2008). AraC regulatory protein mutants with altered effector specificity. Journal of the American Chemical Society, 130(15), 5267–5271. doi:10.1021/ja7109053
Teh, M. Y., Ooi, K. H., Danny Teo, S. X., Bin Mansoor, M. E., Shaun Lim, W. Z., & Tan, M. H. (2019). An expanded synthetic biology toolkit for gene expression control in Acetobacteraceae. ACS Synthetic Biology, 8(4), 708–723. doi:10.1021/acssynbio.8b00168
Teixidó, L., Carrasco, B., Alonso, J. C., Barbé, J., & Campoy, S. (2011). Fur activates the expression of Salmonella enterica pathogenicity island 1 by directly interacting with the hilD operator in vivo and in vitro. PLoS ONE, 6(5). doi:10.1371/journal.pone.0019711
Terpe, K. (2006). Overview of bacterial expression systems for heterologous protein production: From molecular and biochemical fundamentals to commercial systems. Applied Microbiology and Biotechnology, 72(2), 211–222. doi:10.1007/s00253-006- 0465-8
Tobin, J. F., & Schleif, R. F. (1987). Positive regulation of the Escherichia coli L-rhamnose operon is mediated by the products of tandemly repeated regulatory genes. Journal of Molecular Biology, 196(4), 789–799. doi:10.1016/0022-2836(87)90405-0
Tonouchi, N., Sgiyama, M., & Yokozeki, K. (2003). Construction of a vector plasmid for use in Gluconobacter oxydans. Bioscience Biotechnology Biochemistry, 67(1), 211–213. doi:10.1271/bbb.67.211
Venturi, V., Ottevanger, C., Bracke, M., & Weisbeek, P. (1995). Iron regulation of siderophore biosynthesis and transport in Pseudomonas putida WCS358: Involvement of a transcriptional activator and of the Fur protein. Molecular Microbiology, 15(6), 1081– 1093. doi:10.1111/j.1365-2958.1995.tb02283.x
Vía, P., Badía, J., Baldomà, L., Obradors, N., & Aguilar, J. (1996). Transcriptional regulation of the Escherichia coli rhaT gene. Microbiology, 142(7), 1833–1840. doi:10.1099/13500872-142-7-1833
Wickstrum, J. R., & Egan, S. M. (2004). Amino acid contacts between sigma 70 domain 4 and the transcription activators RhaS and RhaR. Journal of Bacteriology, 186(18), 6277– 6285. doi:10.1128/JB.186.18.6277-6285.2004
Wickstrum, J. R., Skredenske, J. M., Balasubramaniam, V., Jones, K., & Egan, S. M. (2010). The AraC/XylS family activator RhaS negatively autoregulates rhaSR expression by preventing cyclic AMP receptor protein activation. Journal of Bacteriology, 192(1), 225– 232. doi:10.1128/JB.00829-08
Wilderman, P. J., Sowa, N. A., FitzGerald, D. J., FitzGerald, P. C., Gottesman, S., Ochsner, U. A., & Vasil, M. L. (2004). Identification of tandem duplicate regulatory small RNAs in Pseudomonas aeruginosa involved in iron homeostasis. Proceedings of the National Academy of Sciences of the United States of America, 101(26), 9792–9797. doi:10.1073/pnas.0403423101
Yamada, Y., & Yukphan, P. (2008). Genera and species in acetic acid bacteria. International Journal of Food Microbiology, 125(1), 15–24. doi:10.1016/j.ijfoodmicro.2007.11.077
Yuan, J., Wu, M., Lin, J., & Yang, L. (2016). Combinatorial metabolic engineering of industrial Gluconobacter oxydans DSM2343 for boosting 5-keto-D-gluconic acid accumulation. BMC Biotechnology, 16(1), 1–14. doi:10.1186/s12896-016-0272-y
Zahid, N., Schweiger, P., Galinski, E., & Deppenmeier, U. (2015). Identification of mannitol as compatible solute in Gluconobacter oxydans. Applied Microbiology and Biotechnology, 5511–5521. doi:10.1007/s00253-015-6626-x
Zhang, L., Lin, J., & Ma, Y. (2010). Construction of a novel shuttle vector for use in Gluconobacter oxydans. Molecular Biotechnology, 46, 227–233. doi:10.1007/s12033- 010-9293-2
Lizenz:In Copyright
Urheberrechtsschutz
Fachbereich / Einrichtung:Mathematisch- Naturwissenschaftliche Fakultät » WE Biologie » Mikrobiologie
Dokument erstellt am:07.03.2022
Dateien geändert am:07.03.2022
Promotionsantrag am:10.11.2021
Datum der Promotion:07.02.2022
english
Benutzer
Status: Gast
Aktionen