• Keine Ergebnisse gefunden

Sulfoquinovose degradation in bacteria

N/A
N/A
Protected

Academic year: 2022

Aktie "Sulfoquinovose degradation in bacteria"

Copied!
116
0
0

Wird geladen.... (Jetzt Volltext ansehen)

Volltext

(1)

Sulfoquinovose degradation in bacteria

Dissertation

Zur Erlangung des akademischen Grades eines Doktors der Naturwissenschaften (Dr. rer. nat.)

an der Universität Konstanz

Mathematisch-Naturwissenschaftliche Sektion Fachbereich Biologie

vorgelegt von

Ann-Katrin Felux

Tag der mündlichen Prüfung: 19. Oktober 2015 1. Referent: PD Dr. David Schleheck

2. Referent: Prof. Dr. Bernhard Schink 3. Referent: Prof. Dr. Valentin Wittmann

Konstanzer Online-Publikations-System (KOPS) URL: http://nbn-resolving.de/urn:nbn:de:bsz:352-0-303851

(2)
(3)
(4)
(5)
(6)

DANKSAGUNG

Zu Beginn möchte ich mich ganz herzlich bei PD Dr. David Schleheck bedanken, für die Möglichkeit, meine Doktorarbeit in seiner Arbeitsgruppe anfertigen zu können, für die hervorragende Betreuung und Unterstützung während der letzten Jahre, für das zahlreiche Korrekturlesen von Manuskripten und der Dissertation sowie darüber hinaus für seine Gastfreundschaft bei diversen Grillabenden.

Vielen Dank an Herrn Prof. Dr. Bernhard Schink für die Übernahme des Koreferates, seine Betreuung innerhalb der Graduiertenschule KoRS-CB und das stete Interesse an meiner Arbeit.

Vielen Dank auch an Herrn Prof. Dr. Valentin Wittmann für die Mitbetreuung im Rahmen der Graduiertenschule KoRS-CB.

Ganz besonders bedanken möchte ich mich bei Michael Weiss für seine große Hilfsbereitschaft und die so wertvollen Diskussionen in allen denkbaren Themengebieten, für die freundschaftliche Zusammenarbeit und dafür, dass ich mich immer auf ihn verlassen konnte, für die zahlreichen verlorenen Wetteinsätze, die meist in Form von Kuchen bezahlt wurden sowie für die vielen privaten Unternehmungen, die, wenn man sie alle aufzählen würde, viele Seiten füllen könnten. Es war eine sehr schöne Zeit!

Vielen Dank auch an meine Bachelorstudentin Laura Rexer für ihre praktische Hilfe im Labor sowie für die abwechslungsreichen und angenehmen 3 Monate, in denen es nie langweilig wurde.

Vielen Dank an Prof. Dr. Alasdair Cook für sein Interesse an meiner Arbeit und zahlreiche wertvolle Diskussionen.

Vielen Dank auch an die gesamte AG Spiteller (Dieter, Kathrin, Ralf, Karin, Daniela) für die freundschaftliche Nachbarschaft in unserem Flügel und die unkomplizierte Mitbenutzung sämtlicher Geräte; an Prof. Dr. Dieter Spiteller für seine große Hilfsbereitschaft bei zahlreichen HPLC-MS Messungen und die stete Diskussionsbereitschaft. Insbesondere möchte ich mich auch bei Karin Denger ganz herzlich bedanken für ihr großes Interesse, die freundschaftliche und angenehme Zusammenarbeit sowie das so gewissenhafte

(7)

Korrekturlesen meiner Arbeit, und Ralf und Kathrin für ihre Gastfreundschaft bei vielen netten Abenden mit Pizza, Crêpes, Chili con Carne, Milchshakes, Kuchen usw.

Vielen herzlichen Dank an die gesamte AG Schink für die unkomplizierte Zusammenarbeit, die Einladung zu ihren Betriebsausflügen und sonstigen Aktivitäten und für die Aufnahme in das Seminarprogramm inkl. Nachsitzungen.

Vielen Dank an die AG van Kleunen für die problemlose Mitbenutzung sämtlicher Räumlichkeiten und Geräte sowie für das angenehme nachbarschaftliche Klima auf dem Stockwerk.

Vielen Dank an die Konstanz Research School Chemical Biology für die Finanzierung meiner Doktorarbeit im Rahmen eines Doktorandenstipendiums sowie für die finanzielle Unterstützung für Reise- und Sachmittel und für die Möglichkeit zur Teilnahme an einer Vielzahl an Kursen, die sehr hilfreich waren.

Vielen herzlichen Dank auch an meine Familie, insbesondere an meine Eltern, die immer an mich geglaubt und mich unermüdlich unterstützt haben. Ohne Euch wäre ich nicht der Mensch geworden, der ich heute bin.

Vielen lieben Dank Sebastian, für deine Liebe und Unterstützung, dass du dein Leben mit mir teilst und immer zu mir stehst.

(8)

CHAPTERS OF THIS THESIS ARE PUBLISHED IN

Denger, K., M. Weiss, A.-K. Felux, A. Schneider, C. Mayer, D. Spiteller, T. Huhn, A. M.

Cook, and D. Schleheck (2014). Sulfoglycolysis in Escherichia coli K-12 closes a gap in the biogeochemical sulfur cycle. Nature 507(7490): 114-117.

Felux, A.-K., P. Franchini, and D. Schleheck (2015). Permanent draft genome sequence of sulfoquinovose-degrading Pseudomonas putida strain SQ1. Standards in Genomic Sciences, 10.1186/s40793-015-0033-x

Felux, A.-K., D. Spiteller, J. Klebensberger, and D. Schleheck (2015). An Entner-Doudoroff pathway for sulfoquinovose degradation in Pseudomonas putida SQ1. Proceedings of the National Academy of Sciences of the U.S.A./pnas.1507049112

FURTHER PUBLICATION NOT INCLUDED IN THIS THESIS

Felux, A.-K., K. Denger, M. Weiss, A. M. Cook and D. Schleheck (2013). Paracoccus denitrificans PD1222 utilizes hypotaurine via transamination followed by spontaneous desulfination to yield acetaldehyde, and finally acetate for growth. J. Bacteriol. 195: 2921- 2930.

(9)
(10)
(11)

TABLE OF CONTENTS

TABLE OF CONTENTS ... I

SUMMARY ... 1

ZUSAMMENFASSUNG ... 3

CHAPTER 1 ... 5

General introduction

CHAPTER 2 ... 13

Sulfoglycolysis in Escherichia coli K-12 closes a gap in the biogeochemical sulfur cycle

CHAPTER 3 ... 35

Permanent draft genome sequence of sulfoquinovose-degrading Pseudomonas putida strain SQ1

CHAPTER 4 ... 47

An Entner-Doudoroff pathway for sulfoquinovose degradation in Pseudomonas putida SQ1

CHAPTER 5 ... 77

General discussion

CHAPTER 6 ... 85

(12)

Appendix

Abbreviations... 85

Record of Contributions ... 91

CHAPTER 7 ... 93

General references

(13)

SUMMARY

The organosulfonate sulfoquinovose (6-deoxy-6-sulfoglucose; SQ) is the polar headgroup of the sulfolipid sulfoquinovosyl diacylglycerol (SQDG), which is present in the photosynthetic membranes of plants, algae, and other phototrophic organisms. Therefore, SQ is highly abundant in nature, where it is degraded by bacteria. Two bacterial model organisms, Escherichia coli K-12 and Pseudomonas putida SQ1 were shown to utilize SQ as carbon and energy source, however, the bacterial degradation pathway(s), and therefore the enzymes and genes involved, remained undefined until now. Since the substrate SQ is available in relevant amounts by chemical synthesis, and also analytical chemistry methods were accessible, the aim of this work was to define the SQ degradation pathways in E. coli K-12 and P. putida SQ1 on the molecular level.

The first SQ degradation pathway was elucidated in E. coli K-12. Differential proteomic and transcriptional analyses revealed a set of inducible candidate genes for a postulated conversion of SQ in analogy to glycolysis (Embden-Meyerhof pathway) for glucose-6- phosphate. This “sulfoglycolysis” pathway was confirmed to involve four newly discovered enzymes, which were heterologously produced and purified, and three newly discovered organosulfonate intermediates, which were identified by HPLC-mass spectrometry: SQ is isomerized to 6-deoxy-6-sulfofructose (SF) by a SQ isomerase (YihS), SF phosphorylated to 6-deoxy-6-sulfofructose-1-phosphate (SFP) by a SF kinase (YihV), and SFP is cleaved into dihydroxyacetonephosphate (DHAP) and 3-sulfolactaldheyde (SLA) by a SFP aldolase (YihT). The DHAP powers energy conservation and growth of the bacterium while the SLA is reduced to 2,3-dihydroxypropane-1-sulfonate (DHPS) by a SLA reductase (YihU), and excreted, due to missing desulfonation genes in E. coli K-12. The DHPS however can be degraded completely by other bacteria, e.g. Cupriavidus pinatubonensis JMP134, as demonstrated in this study. Thus, it was shown that SQ can be mineralized within a bacterial community. Sulfoglycolysis is encoded in a ten-gene cluster (inclusive candidate genes for transport and regulation) that is found in almost all available E. coli genomes (>91%) and in many other Enterobacteria. These observations suggest a significant role of sulfoglycolysis in the alimentary tract of all omnivores and herbivores, and in plant pathogens.

The second SQ degradation pathway was uncovered in P. putida strain SQ1, whose genome was sequenced and annotated in this study. A set of inducible candidate genes for a postulated conversion of SQ analogous to the Entner-Doudoroff pathway for glucose-6-phosphate was

(14)

revealed by proteomic and transcriptional analyses. This Entner-Doudoroff pathway for SQ was confirmed to involve five newly discovered enzymes, which were also heterologously produced and purified, and three newly discovered organosulfonate intermediates, which were also identified by HPLC-mass spectrometry: SQ is oxidized to 6-deoxy-6- sulfogluconolactone (SGL) by a SQ dehydrogenase, SGL hydrated to 6-deoxy-6- sulfogluconate (SG) by a SGL lactonase, SG dehydrated to 2-keto-3,6-dideoxy-6- sulfogluconate (KDSG) by a SG dehydratase, and the KDSG is cleaved into pyruvate and SLA by a KDSG aldolase. The pyruvate powers energy conservation and growth of the bacterium while the SLA is oxidized to 3-sulfolactate (SL) by a SLA dehydrogenase, and excreted; thus, also P. putida SQ1 is unable to catalyze a desulfonation. However, the SL can be degraded completely by other bacteria, e.g., by Paracoccus pantotrophus NKNCYSA, as demonstrated in previous studies, and also this time, a bacterial community is necessary to fully mineralize SQ. The SQ Entner-Doudoroff pathway in P. putida SQ1 is encoded in a thirteen-gene cluster (inclusive candidate genes for, e.g., transport and regulation), and homologous gene clusters were found to be widely distributed among typical marine, freshwater and soil Proteobacteria, indicating that SQ utilization is a widespread and important, but yet under-recognized trait of bacteria in all environments where SQ is produced and degraded.

(15)

ZUSAMMENFASSUNG

Das Organosulfonat Sulfoquinovose (6-Desoxy-6-sulfoglukose; SQ) ist die polare Kopfgruppe des Sulfolipids Sulfoquinovosyl diacylglycerol (SQDG), das in den photosynthetischen Membranen von Pflanzen, Algen und anderen phototropen Organismen und somit in großen Mengen in der Natur vorkommt und dort von Bakterien wieder abgebaut wird. Es wurde bereits gezeigt, dass zwei bakterielle Modelorganismen, Escherichia coli K- 12 und Pseudomonas putida SQ1 SQ als Kohlenstoff- und Energiequelle nutzen können, der bakterielle Abbauweg sowie die beteiligten Enzyme und Gene blieben bislang jedoch unbekannt. Da das Substrat SQ nun durch chemische Synthese in ausreichender Menge zur Verfügung steht und zusätzlich chemisch-analytische Methoden zugänglich sind, war nun das Ziel dieser Arbeit, die SQ-Abbauwege in E. coli K-12 und P. putida SQ1 auf molekularer Ebene aufzuklären.

Der erste Abbauweg für SQ wurde in E. coli K-12 aufgeklärt. Durch differenzielle proteomische und transkriptionelle Analysen wurde eine Reihe von induzierbaren Genen für einen SQ-Abbau, in Analogie zur Glykolyse (Embden-Meyerhof-Weg) für Glukose-6- phosphat, postuliert. Es wurde bestätigt, dass für die Sulfoglykolyse vier neue Enzyme verantwortlich sind, die heterolog produziert und gereinigt wurden. Zusätzlich wurden mittels HPLC-Massenspektrometrie drei neue Organosulfonate als Intermediate identifiziert: SQ wird von einer SQ-Isomerase (YihS) zu 6-Desoxy-6-sulfofruktose (SF) isomerisiert, SF von einer SF Kinase (YihV) zu 6-Desoxy-6-sulfofruktose-1-phosphat (SFP) phosphoryliert und SFP von einer SFP Aldolase (YihT) zu 4-Sulfolaktaldehyd (SLA) und Dihydroxyacetonphosphat (DHAP) gespalten. DHAP geht in den zentralen Stoffwechsel der Zelle ein und ist für das bakterielle Wachstum verantwortlich. SLA wird von einer SLA Reduktase (YihU) zu 2,3- Dihydroxypropan-1-sulfonat reduziert (DHPS) und ausgeschieden. E. coli K-12 besitzt keine Gene für eine Desulfonierung, allerdings sind andere Bakterien, wie z.B. Cupriavidus pinatubonensis JMP 134 in der Lage, DHPS vollständig abzubauen. Somit wurde gezeigt, dass eine bakterielle Gemeinschaft für die Mineralisierung von SQ notwendig ist. Die Sulfoglykolyse ist in einem Zehn-Gen Cluster kodiert (inklusive vorhergesagter Gene für Transport und Regulation), das in nahezu allen momentan verfügbaren E. coli Genomen (>91%) und in vielen weiteren Enterobakterien zu finden ist. Diese Beobachtungen deuten auf eine bedeutende Rolle der Sulfoglykolyse im Verdauungstrakt aller omni- und herbivoren Lebewesen sowie in Pflanzenpathogenen hin.

(16)

Der zweite Abbauweg für SQ wurde in P. putida SQ1 aufgeklärt, dessen Genom in dieser Arbeit sequenziert und analysiert wurde. Durch differenzielle proteomische und transkriptionelle Analysen wurden eine Reihe von induzierbaren Genen für einen SQ-Abbau, dieses Mal in Analogie zum Entner-Doudoroff-Weg für Glukose-6-phosphat, postuliert. Es wurde bestätigt, dass für den SQ Entner-Doudoroff-Weg fünf neue Enzyme verantwortlich sind, die heterolog produziert und gereinigt wurden. Weiterhin wurden mit Hilfe von HPLC- Massenspektrometrie drei neue Organosulfonate als Intermediate identifiziert: SQ wird von einer SQ Dehydrogenase zu 6-Desoxy-6-sulfoglukonolakton (SGL) oxidiert, SGL von einer SGL Laktonase zu 6-Desoxy-6-sulfoglukonat (SG) hydratisiert, SG von einer SG Dehydratase zu 2-Keto-3,6-didesoxy-6-sulfoglukonat (KDSG) dehydratisiert und KDSG von einer KDSG Aldolase zu Pyruvat und SLA gespalten. Pyruvat geht in den zentralen Stoffwechsel der Zelle ein und ist für das bakterielle Wachstum verantwortlich. SLA wird von einer SLA Dehydrogenase zu 3-Sulfolaktat (SL) oxidiert und ausgeschieden, denn auch P.

putida SQ1 besitzt keine Gene für eine Desulfonierung. Es wurde jedoch in früheren Studien gezeigt, dass andere Bakterien, wie z.B. Paracoccus pantotrophus NKNCYSA, SL komplett abbauen können. Daher ist auch in diesem Fall eine bakterielle Gemeinschaft für eine vollständige Mineralisierung von SQ notwendig. Der SQ Entner-Doudoroff Weg in P. putida SQ1 wird von einem Dreizehn-Gen Cluster (inklusive vorhergesagter Gene für z.B. Transport und Regulation) kodiert und homologe Gencluster sind unter einer Vielzahl von typischen marinen, Süßwasser und Boden Proteobakterien weit verbreitet. Diese Beobachtung lässt darauf schließen, dass die Verwertung von SQ weit verbreitet, wichtig und eine bis jetzt unterschätzte Eigenschaft von Bakterien in allen Lebenswelten ist, in denen SQ produziert und wieder abgebaut wird.

(17)

CHAPTER 1

General introduction

The biogeochemical sulfur cycle in nature

Sulfur is a ubiquitous element and sulfur containing compounds are essential for all life, e.g.

in form of amino acids, vitamins and cofactors. The sulfur entry into the atmosphere can occur not only naturally by volcanic eruptions, but also industrially by the burning of fossil fuels, thus the sulfur entry is predominantly in form of sulfur dioxide (SO2), but also in form of hydrogen sulfide (H2S) (Guo et al. 2004). Subsequently the sulfur containing compounds are carried back to the earth surface with rain and snow, where they can, for example, be oxidized by microbial activities to sulfate. Inorganic sulfate is the predominant external source of sulfur for e.g. plants in soils, where it can be taken up by the roots and used by the plant for the biosynthesis of essential organo sulfur compounds on the one hand (e.g.

Rennberg 1984) or used by sulfate reducing bacteria as terminal electron acceptor on the other hand (e.g. Postgate 1965, Hamilton 1985). Due to an increasing diminution of sulfur deposition in rainfall concomitant with an intensive agriculture nowadays, many soils are becoming sulfur deficient (Roy et al. 2003). Therefore, a detailed understanding of the effective re-mobilization and recycling of the organically bound sulfur through microbial biodegradation has become even more important, in order to sustain the sulfur cycle in our ecosystem (Harwood and Nicholls 1979).

Organosulfonates – an overview

Organosulfonates are a class of organo sulfur compounds that contribute substantially to the sulfur cycle in nature. They either can occur naturally through biosynthesis (Shibuya et al.

1963, Huxtable 1986, Autry and Fitzgerald 1990, Vairavamurthy et al. 1994), or they can occur xenobiotically from industrial chemical synthesis (Cook et al. 2007). A highly abundant, naturally occurring organosulfonate is the sulfo-sugar sulfoquinovose (SQ), which is the topic of this thesis. An example of a relevant organosulfonate of industrial origin is the widely used laundry surfactant linear-alkylbenzenesulfonate (LAS) (e.g. Cook 1998). The microbial degradation of organosulfonates represent a major component of both the biological sulfur and carbon cycle (Harwood and Nicholls 1979), in which, for example, heterotrophic

(18)

bacteria can utilize organosulfonates as a carbon and energy source concomitant with the liberation the sulfonate-sulfur as inorganic sulfate (e.g. Cook and Denger 2002).

Organosulfonates are characterized by a sulfonic acid group (-SO3H) that is directly linked to a carbon atom (R-SO3H), which is strongly acidic and hence, negatively charged over the whole physiological pH range (R-SO3-

). Therefore organosulfonates require active transport across biological membranes for their degradation (Huxtable 1992, Cook et al. 1999). This class of organo sulfur compounds differs fundamentally from other compounds containing a sulfonate moiety, e.g. sulfate esters, where the sulfur atom is linked to oxygen (O-sulfonates) or N-sulfonates, where the sulfur atom is linked to a nitrogen atom. The latter can be easily hydrolyzed, whereas organosulfonates were reported to be highly stable against chemical hydrolysis (Wagner and Reid 1931) and therefore, far more sophisticated desulfonation strategies are needed (e.g. Dodgson and White 1983). In this context, bacteria and fungi play a remarkable role, since these microorganisms – to our current knowledge – are the only ones that are able to effectively cleave this stable carbon-sulfur bond (Huxtable 1992).

Currently, four different desulfonation mechanisms including the respective enzymes are known for aliphatic organosulfonates: desulfonating monooxygenases of alkanesulfonates (Thysse and Wanders 1974), a sulfoacetaldehyde acetyltransferase (Xsc), which converts sulfoacetaldehyde to acetyl phosphate and sulfite (Ruff et al. 2003), a 3-sulfolactate sulfo- lyase (SuyAB), which converts 3-sulfolactate to pyruvate and sulfite (Rein et al. 2005), and a cysteate sulfo-lyase (CuyA), which converts cysteate to pyruvate, ammonium and sulfite (Denger et al. 2006). With desulfonation, the toxic product sulfite is formed, and bacteria protect themselves from its cytotoxic effects by expressing specific sulfite exporters (e.g.

Felux et al. 2013) or by oxidizing sulfite to sulfate with sulfite dehydrogenase enzymes (e.g.

Kappler et al. 2000, Denger et al. 2011)

Notably, there are also bacteria that are unable to abstract the sulfonate moiety from organosulfonates, but utilize only parts of the compounds for growth and excrete the remainder organosulfonate-intermediates. These intermediates can then be fully dissimilated by other bacteria to CO2 and sulfite/sulfate. Thus, for some organosulfonates, a bacterial community is required for the complete degradation and recycling of the sulfur (e.g.

Schleheck et al. 2000, Dong et al. 2004, Schleheck et al. 2004, Denger et al. 2012, Weiss et al. 2012).

(19)

Sulfoquinovose – a naturally occurring organosulfonate

Sulfoquinovose (6-deoxy-6-sulfoglucose; SQ) is the polar headgroup of the plant sulfolipid sulfoquinovosyl diacylglycerol (1,2-diacyl-3-sulfoquinovosylglycerol; SQDG) and thus, a naturally occurring organosulfonate.

SQDG is located in the photosynthetic membranes of all higher plants, mosses, ferns, algae and most photosynthetic bacteria as well as in some non-photosynthetic bacteria (Harwood and Nicholls 1979, Benning et al. 1993, Benning 1998). With an estimated annual production of 3.6 x 1010 tons of SQDG, it is thought to be one of the most abundant sulfur-organic compounds in the biotic world following methionine, glutathione and cysteine and thus, makes a major contribution to the global sulfur cycle in nature (Harwood and Nicholls 1979).

However, the biosynthesis pathway, the function of the sulfolipid in plants, as well as the biodegradation pathway for SQDG and SQ is not fully understood until now (Pugh et al.

1995, Benning 1998, Yu et al. 2002, Benning 2007, Zolghadr et al. 2015). Nevertheless it is thought that a complete catabolism of SQ concomitant with a release of sulfate is dominated by soil microorganisms (Martelli and Benson 1964, Strickland and Fitzgerald 1983).

Due to the fact that SQ is highly abundant, the degradation, and therefore the recycling of sulfur is a crucial step in sustaining the sulfur cycle in nature (Roy et al. 2003, Denger et al.

2012, Denger et al. 2014, Felux et al. 2015A, Felux et al. 2015B). Only few years after the discovery of SQ, the first bacteria were described to degrade SQ quantitatively to sulfate via intracellular cysteate and sulfoacetate (Martelli and Benson 1964, Martelli and Souza 1970), however, these organisms were not further analyzed for their pathways, and were not maintained in any culture collection (Cook and Denger 2002). More than ten years ago it was shown that five different bacterial soil isolates (Pseudomonas, Agrobacterium and Klebsiella species) were able to degrade SQ, and these bacteria excreted substoichiometric amounts of sulfate, 3-sulfolactate (SL) or 2,3-dihydroxypropane-1-sulfonate (DHPS) into the growth medium (Roy et al. 2000, Roy et al. 2003). Three years ago it was demonstrated, for the first time in pure cultures, that sulfoquinovose can be fully degraded to sulfate in defined sets of two-member bacterial communities (Figure 1): In a first community, Klebsiella oxytoca TauN1 utilizes SQ and excretes DHPS in a first tier, and Cupriavidus pinatubonensis JMP134 degrades the DHPS completely to sulfate in the second tier. In a second model community, Pseudomonas putida SQ1 utilizes SQ and excretes SL, which is completely degraded by Paracoccus pantotrophus NKNCYSA (Denger et al. 2012).

Degradation pathways for the second tier, i.e., for the complete degradation of SL and DHPS to sulfate, are known since recently (Rein et al. 2005, Denger and Cook 2010, Mayer et al.

(20)

2010): In a bifurcated pathway, SL is oxidized to sulfopyruvate, and then, the sulfopyruvate is either aminated to cysteate, and the cysteate is desulfonated to pyruvate as catalyzed by cysteate sulfo-lyase A (CuyA; EC 4.4.1.25); or the sulfopyruvate is decarboxylated to sulfoacetaldehyde, which is subsequently desulfonated to acetyl phosphate, as catalyzed by sulfoacetaldehyde acetyltransferase (Xsc; EC 2.3.3.15) (Denger et al. 2009). Additionally, a third option was found, a direct desulfonation of SL to pyruvate, as catalyzed by sulfolactate sulfo-lyase AB (SuyAB; EC 4.4.1.24) (Rein et al. 2005). DHPS was shown to be oxidized to SL by NAD(P)+-coupled DHPS-dehydrogenase (HpsN; EC 1.1.1.308) and subsequently channeled into one of the three pathways for desulfonation of SL (Denger and Cook 2010, Mayer et al. 2010). However, no demonstration of any of the proposed pathways for SQ degradation, i.e. from SQ to the C3-sulfonated intermediates, was provided on the molecular level until today.

Figure 1. Two pure-culture laboratory model systems for a complete degradation of SQ.

Pseudomonas putida SQ1 and Klebsiella oxytoca TauN1 each are able to utilize SQ and excrete sulfonated C3 intermediates into the growth medium, 3-sulfolactate or 2,3-dihydroxypropane-1- sulfonate (DHPS), respectively, which are mineralized by bacterial strains of the second tier, Paracoccus pantotrophus NKNCYSA or Cupriavidus pinatubonensis JMP134, respectively (adapted from Denger et al. 2012).

Due to the structural similarity of SQ to glucose-6-phosphate (Figure 2A and B), it seems plausible that the mechanism for degradation of SQ could proceed in analogy to glucose/glucose-6-phosphate degradation (Roy et al. 2003). In this case, there are three options for proposed pathways: One of the two major pathways would proceed in analogy to the Embden-Meyerhof-Parnas pathway (glycolysis), where glucose is phosphorylated to

SO42-

SO42-

Sulfoquinovose (SQ)

Growth of Pseudomonas putidaSQ1

Growth of

Paracoccus panotrophusNKNCYSA

Growth of Klebsiella oxytocaTauN1

Growth of

Cupriavidus pinatubonensisJMP134 3-sulfolactate

(SL)

2,3-dihydroxypropane-1-sulfonate (DHPS)

(21)

glucose-6-phosphate, isomerized to fructose-6-phosphate, phosphorylated to fructose-1,6- bisphosphate and cleaved into dihydroxyacetone phosphate (DHAP) and glycerol aldehyde-3- phosphate (GAP). The latter is converted into pyruvate in the lower glycolysis pathway. The second major route proceeds in analogy to the Entner-Doudoroff pathway for glucose-6- phosphate, which is initially oxidized to 6-phosphogluconolactone, then hydrated to 6- phosphogluconate, dehydrated to the key intermediate 2-keto-3-deoxy-6-phosphogluconate (KDPG) and cleaved into pyruvate and GAP; the latter is converted into pyruvate in the lower glycolysis pathway. The third optional pathway could operate via a pentose-phosphate-type pathway, which is rather an anabolic than a catabolic pathway, since the important reducing equivalent for biosynthetic processes, NADPH, is formed concomitant with ribulose-5- phosphate, which serves as a building block for the synthesis of e.g., DNA, ATP, NADH or Coenzyme A.

Figure 2. Chemical structure of (A) the naturally occurring organosulfonate sulfoquinovose (SQ) in comparison to (B) glucose-6-phosphate.

The plant sulfolipid sulfoquinovosyl diacylglycerol (SQDG) - function and biosynthesis

The plant sulfolipid SQDG was discovered during 35S-labelling experiments of different photosynthetic bacteria, algae and plants over half a century ago, in which a sulfur-containing compound was uncovered in the lipid phase of cell extracts (Benson et al. 1959, Benson 1963). Since SQDG is widely distributed in photosynthetic membranes, it was first supposed to play a crucial role during photosynthesis (Barber and Gounaris 1986); but later on it was shown that SQDG-deficient mutants of oxygenic (cyanobacterium Synechococcus) and anoxygenic photosynthetic organisms (Rhodobacter sphaeroides) and also of higher plants (Arabidopsis) were not impaired in growth or electron transport under normal growth conditions (Benning et al. 1993, Güler et al. 1996, Benning 1998, Essigmann et al. 1998, Yu et al. 2002, Yu and Benning 2003, Aoki et al. 2004). Further investigations on the function

B

A

(22)

and role of SQDG revealed that under phosphate-limited growth conditions the phospholipid phosphatidylglycerol (PG) can be replaced by SQDG in photosynthetic bacteria, algae and plants, which seems to be a reasonable strategy of photosynthetic organisms to internally recycle the phosphorous bound in PG-lipids (e.g. Benning et al. 1993, Güler et al. 1996, Essigmann et al. 1998, Yu et al. 2002, Aoki et al. 2004). These results suggest that SQDG is not essential for photosynthesis. However it is thought to play a role in regulating the activity of photosystem II and to substitute PG under limited phosphate availability to maintain the proper balance of anionic charge in the thylakoid membranes (Mizusawa and Wada 2012) and therefore for accurate protein import into chloroplasts.

About thirty years after its discovery, first indirect evidence for a biosynthesis pathway for SQDG was provided, where the sulfonate moiety is introduced at the level of hexose (Heinz et al. 1989, Benning and Somerville 1992A, Benning and Somerville 1992B); this resulted 1995 in the formulation of the so-called “sugar-nucleotide-pathway hypothesis” (Pugh et al.

1995). In this hypothesis, SQDG biosynthesis is a two steps mechanism: First, uridine diphosphate (UDP)-SQ is formed from UDP-glucose and sulfite and second, UDP-SQ is fused with a diacylglycerol (DAG) molecule forming SQDG while the UDP is recovered (Figure 3).

Figure 3. The biosynthetic pathway for sulfoquinovosyl diacylglycerol (SQDG). UDP-sulfoquinovose is formed from UDP-glucose and linked to a diacyl glycerol (DAG), which yields SQDG. R1 and R2 indicate acyl chains with variable length and degree of saturation. SQDG, sulfoquinovosyl diacylglycerol; UGP3, UDP-glucose pyrophosphorylase; SQD1, UDP-sulfoquinovose synthase;

SQD2, SQDG synthase. (adapted from Shimojima 2011).

(23)

However, only in the last two decades substantial progress was made in understanding the biosynthesis and exact mechanism how sulfite is incorporated into UDP-glucose, based on the sugar nucleotide pathway hypothesis from Pugh et al, including the genes in cyanobacteria and plants (e.g. Güler et al. 2000, Sanda et al. 2001, Yu et al. 2002, Okazaki et al. 2009). To briefly summarize all these results, it is now clear that SQDG assembly mainly requires UDP- glucose pyrophosphorylase (UGP3), UDP-sulfoquinovose (UDP-SQ) synthase (SQD1) and SQDG synthase (SQD2) (Figure 3): UGP3 produces UDP-glucose (from glucose-1- phosphate), which is the substrate for UDP-SQ synthase to form UDP-SQ. The latter is transferred to a diacylglycerol residue by the activity of SQDG synthase to produce SQDG (Essigmann et al. 1998, Sanda et al. 2001, Yu et al. 2002, Okazaki et al. 2009). Notably, only very recently, a more detailed reaction mechanism including the corresponding enzyme for the formation of UDP-SQ in the hyperthermophilic archaeon Sulfolobus acidocaldarius was proposed (Zolghadr et al. 2015) as catalyzed by a UDP-SQ synthase, which belongs to the short-chain-dehydrogenase/reductase superfamily of enzymes (Field and Naismith 2003, Kavanagh et al. 2008, Zolghadr et al. 2015).

Aims of this study

At the beginning of the experimental work for this thesis, it was known that Escherichia coli K-12 substrain MG1665 is able to utilize SQ as a sole source of carbon and energy for growth, and the first candidate genes of its SQ degradation pathway had been proposed through differential proteomic using the known genome sequence of E. coli K-12. A second bacterium that degrades SQ was also available, Pseudomonas putida strain SQ1 (Denger et al.

2012), but not its genome sequence. In this study, the genome sequence of P. putida SQ1 should be established and used for differential proteomic and transcriptomic analyses, in order to reveal candidate genes also for a second SQ degradation pathway. Then, a functional confirmation of the role of the candidate genes in the respective SQ degradation pathways of E. coli K-12 and of P. putida SQ1, should be attempted, e.g., by cloning and heterologous expression of the candidate genes, and functional testing of the purified proteins for their attributed enzyme activity

(24)
(25)

CHAPTER 2

Sulfoglycolysis in Escherichia coli K-12 closes a gap in the biogeochemical sulfur cycle

Karin Denger1, Michael Weiss2, Ann-Katrin Felux2, Alexander Schneider3, Christoph Mayer3, Dieter Spiteller1, Thomas Huhn4, Alasdair M. Cook1 & David Schleheck1

1Department of Biology, University of Konstanz, Germany

2Konstanz Research School Chemical Biology, University of Konstanz, Germany

3Interfaculty Institute of Microbiology and Infection Medicine, University of Tübingen, Germany

4Department of Chemistry, University of Konstanz, Germany

This chapter was published in Nature. 2014, 507(7490):114-117.

(26)

ABSTRACT

Sulfoquinovose (SQ, 6-deoxy-6-sulfoglucose) has been known for 50 years as the polar headgroup of the plant sulfolipid (Benson 1963, Benning 2007) in the photosynthetic membranes of all higher plants, mosses, ferns, algae, and most photosynthetic bacteria (Benning 1998). It is also found in some non-photosynthetic bacteria (Harwood and Nicholls 1979), and SQ is part of the surface layer of some Archaea (Meyer et al. 2011). The estimated annual production of SQ (Harwood and Nicholls 1979) is 10,000,000,000 tones (10 petagrams), thus it comprises a major portion of the organo-sulfur in nature, where SQ is degraded by bacteria (Martelli 1967, Roy et al. 2003). However, despite evidence for at least three different degradative pathways in bacteria (Martelli 1967, Roy et al. 2003, Denger et al.

2012), no enzymic reaction or gene in any pathway has been defined, although a sulfoglycolytic pathway has been proposed (Roy et al. 2003).

Here we show that Escherichia coli K-12, the most widely studied prokaryotic model organism, performs sulfoglycolysis, in addition to standard glycolysis. SQ is catabolized through four newly discovered reactions that we established using purified, heterologously expressed enzymes: SQ isomerase, 6-deoxy-6-sulfofructose (SF) kinase, 6-deoxy-6- sulfofructose-1-phosphate (SFP) aldolase, and 3-sulfolactaldehyde (SLA) reductase. The enzymes are encoded in a ten-gene cluster, which probably also encodes regulation, transport and degradation of the whole sulfolipid; the gene cluster is present in almost all (>91%) available E. coli genomes, and is widespread in Enterobacteriaceae. The pathway yields dihydroxyacetone phosphate (DHAP), which powers energy conservation and growth of E. coli, and the sulfonate product 2,3-dihydroxypropane-1-sulfonate (DHPS), which is excreted. DHPS is mineralized by other bacteria, thus closing the sulfur cycle within a bacterial community.

INTRODUCTION, RESULTS, AND DISCUSSION

Recent work showed that environmental isolates of Klebsiella spp. (Enterobacteriaceae) convert SQ quantitatively to DHPS (Roy et al. 2003, Denger et al. 2012), and we hypothesized that utilization of SQ might be a property of Enterobacteriaceae. We found that four genome-sequenced E. coli K-12 substrains (BW25113, DH1, MG1655 and W3100), after subculturing, grew with SQ within 1 to 3 days. We chose to work (largely) with the fastest-growing substrain, MG1655. The organism used SQ as a sole source of carbon and energy with a molar-growth yield of 3 g of protein per mol of SQ carbon, whereas glucose gave about 6 g of protein per mol of carbon; the latter value represented mass balance of

(27)

carbon as biomass and CO2 (Cook 1987). However, approximately 1 mol of DHPS per mol of SQ was released into the growth medium (Figure 1a), as observed with Klebsiella oxytoca (Denger et al. 2012). Thus, there was complete mass balance for carbon and for sulfur from SQ. The growth rate with SQ was 0.13 h-1 (0.5 h-1 with glucose), and the specific degradation rate for SQ in vivo was 120 mU per mg of protein (1 mU = 1 nmol min-1). We concluded that SQ is metabolized to a C3 sulfonate, which is excreted as DHPS, and that the remainder of the molecule is utilized for growth (Figure 2a).

The out-grown culture was filter-sterilized and inoculated with Cupriavidus pinatubonensis JMP134, which can utilize DHPS for growth (Mayer et al. 2010), but cannot utilize SQ (Denger et al. 2012). C. pinatubonensis grew with the DHPS formed from SQ by E. coli, and released its sulfonate-sulfur quantitatively as sulfate (Figure 1b) using a pathway described elsewhere (Mayer et al. 2010). We thus demonstrated mineralization of SQ in a laboratory model system.

Figure 1. Complete degradation of sulfoquinovose during growth. a, Growth of E. coli K-12 substrain MG1655 with SQ and excretion of 2,3-dihydroxypropane-1-sulfonate (DHPS). b, Growth of C.

pinatubonensis JMP134 with the DHPS formed from SQ by E. coli. Data from representative growth experiments (n = 3) are shown. To allow a compact graph, sulfate release and not total sulfate is shown.

(28)

Proteins from whole cells of E. coli K-12 grown with glucose or SQ were subjected to two dimensional polyacrylamide gel electrophoresis (2D-PAGE) (Extended Data Figure 1) and examined by peptide fingerprinting-mass spectrometry (PF-MS) (Extended Data Table 1).

The immediately relevant, apparently SQ-inducible proteins (see Extended Data Figure 1 and Extended Data Table 1) were attributed to b3878 (also known as yihQ (b numbers are locus tags); predicted to be an -glucosidase), b3879 (also known as yihR; predicted to be an epimerase), b3880 (yihS; predicted to be an isomerase), b3881 (yihT; predicted to be an aldolase), and b3882 (yihU; predicted to be an NAD+/NADH-linked dehydrogenase/reductase). Transcriptional analyses for the gene cluster b3879-b3882, as well as for b3883 (also known as yihV; predicted to be a sugar kinase), confirmed a strong inducible transcription during growth with SQ, but not during growth with glucose (Extended Data Figure 2). Furthermore, single-gene knockouts (in substrain BW25113; Baba et al.

2006) in genes b3876 (also known as yihO; predicted to be a major facilitator superfamily (MFS)-type transporter), b3880, b3881 and b3883 did not grow with SQ, which confirmed and expanded on the proteomic and transcriptional data (Figure 2b).

We thus identified a gene cluster in E. coli K-12 that contained SQ-inducible, essential genes for catabolism of SQ, but we still did not know which pathway was involved. A sulfoglycolytic pathway would involve a hypothetical 3-sulfolactaldehyde (SLA) reductase to yield DHPS in the final reaction (apart from export) (Figures 1a and 2a), whereas a hypothetical SQ dehydrogenase as the first reaction would lead into hypothetical Entner- Doudoroff-type or pentose-phosphate-type pathways, or another novel pathway. An SLA- reductase was detected (assayed as DHPS oxidation) in cell-free extracts of SQ-grown substrain MG1655 at a specific activity of 420 mU per mg of protein, which exceeds the specific degradation rate for SQ in vivo and, thus, was sufficient to explain growth. This enzyme activity was not detected in extracts of glucose-grown cells. The enzyme was, thus, confirmed to be inducible, and it was specific for NAD+; NADP+ was not a substrate.

Furthermore, SQ did not lead to reduction of NAD+ or of NADP+ in the extracts of SQ- or glucose-grown cells, hence, hypothetical SQ dehydrogenase was not detectable. These data led us to predict the sulfoglycolytic pathway depicted in Figure 2a, including the requirement for sulfonate import and export across the cell membrane (Graham et al. 2002, Mampel et al.

2004, Mayer and Cook 2009).

(29)

Figure 2. The four core enzyme reactions of sulfoglycolysis, with transport, and the corresponding genes in a ten-gene cluster in E. coli K-12. a, SQ is metabolized by four enzymes (shown in color) to a C3 sulfonate, DHPS, which is excreted, and the remainder of the molecule is used for growth. For comparison, the analogous enzyme reactions for the catabolism of (unsubstituted) glucose through the glycolytic pathway in E. coli, are also shown (dashed arrows). Fba, fructose bisphosphate aldolase;

GAP, glyceraldehyde-3-phosphate; Pfk, phosphofructokinase; Pgi, phosphoglucose isomerase; PTS, phosphotransferase system permease; Tpi, triose phosphate isomerase. b, The EcoGene E. coli website (http://www.EcoGene.org) uses the abbreviation ‘yih’ for most of these genes; we have retained this nomenclature. Vertical stripes, genes confirmed as being essential for growth with SQ by mutational analysis; horizontal stripes, genes confirmed as being inducible for growth with SQ by proteomic and/or transcriptional analyses; box-framed genes, genes encoding the four core enzymes of the pathway (shown in a) that were subject of heterologous expression, purification, and functional characterization.

The four predicted core enzymes of the pathway (Figure 2a) were heterologously expressed and purified as His-tagged proteins, b3880 (putative isomerase), b3883 (putative sugar kinase), b3881 (putative aldolase) and b3882 (putative reductase) (Extended Data Figure 3).

Protein b3882 was shown to encode SLA reductase. First, we partially purified and identified (PF-MS) the wild type enzyme in cell-free extracts of substrain MG1655 (see above), and second, we examined the recombinant protein (see below). In both cases, we identified that b3882 represents an SLA reductase; the enzyme showed no activity with 4-hydroxybutyrate (Saito et al. 2009).

The heterologously expressed and purified putative isomerase (b3880) caused about one- fourth of the SQ in the reaction mixture to disappear, as observed by high-pressure liquid chromatography-mass spectrometry (HPLC-MS), and a new peak was formed that eluted with shorter retention time, but exhibited the same relative mass (Mr = 244 Dalton (Da); observed

(30)

as a quasi-molecular ion in the negative ion mode ([M-H]-) at a mass-to-charge ratio (m/z) of 243) (Figure 3a, b). The new peak was confirmed to represent 6-deoxy-6-sulfofructose (SF), as proposed elsewhere (Roy et al. 2003), by the HPLC separation pattern (Extended Data Figure 4), by the matching exact mass of the [M-H]- ion (Extended Data Figure 4), and by its MS/MS fragmentation pattern (Extended Data Figure 5). Thus we confirmed that b3880 catalyzed the SQ isomerase reaction.

The reaction mixture was augmented with ATP and the putative sugar kinase (b3883). The peaks of SQ and SF partially disappeared and a new peak was formed (Figure 3c). This new peak was confirmed to represent 6-deoxy-6-sulfofructose-1-phosphate (SFP), as proposed elsewhere (Roy et al 2003), by the matching exact mass of the [M-H]- ion (observed mass, 322.9877 Da; theoretical mass of C6H12O11PS-, 322.9843 Da) and by its fragmentation pattern (Extended Data Figure 6). HPLC confirmed that ATP disappeared and ADP was formed during the reaction and, furthermore, that SFP was converted back to SF when alkaline phosphatase was added to a preparation of SFP (not shown). Thus, with b3883, we expressed an ATP-dependent kinase that phosphorylated SF to SFP.

The reaction mixture was augmented with the putative aldolase (b3881). The peak for SFP partially disappeared, and two new peaks were formed (Figure 3d). The first new peak was identified to represent DHAP, as proposed elsewhere (Roy et al. 2003), with an authentic DHAP standard. The second new peak was confirmed to represent SLA, as proposed elsewhere (Roy et al. 2003), by the matching mass of the [M-H]- ion (Mr = 154; observed as [M-H]- ion at m/z = 153) and by its fragmentation pattern (Extended Data Figure 7); the same peak was observed when we used recombinant SLA reductase in reverse to oxidize DHPS to SLA (see above). Thus, with b3881, we expressed an aldolase that cleaved SFP into DHAP and SLA. The SFP turnover was incomplete (Figure 3d); the equilibrium of the corresponding enzyme reaction in glycolysis (fructose-1,6-bisphosphate aldolase) lies far to the left (Cornish-Bowden 1981), that is, hardly any products are formed.

However, when NADH and the recombinant SLA-reductase (b3882) were added, the peak for SFP was further diminished, as was the peak for SQ, and that for the aldolase-reaction product DHAP, was further increased (Figure 3e). In addition, the peak for SLA had disappeared, and the peak for the anticipated sulfonate product, DHPS, was formed (Figure 3e and Extended Data Figure 7). After an extended incubation of the four-enzyme reaction (see Figure 3e, f), the peaks for SQ and SFP had almost completely disappeared, and the peaks for DHAP and DHPS, had further increased.

(31)

Figure 3. Illustration of the reactions of the four core enzymes of sulfoglycolysis in vitro. The transformation of SQ to SF, SFP, DHAP and SLA, and DHPS, by successive addition of recombinantly expressed pathway enzymes was followed by HPLC-ESI-MS. a, Sample of SQ in reaction buffer (t = 0 min). b, Sample after addition of isomerase (b3880) (t = 60 min). c, Sample after addition of ATP and kinase (b3883) (t = 120 min). d, Sample after addition of aldolase (b3881) (t = 180 min). e, Sample after addition of NADH and reductase (b3882) (t = 240 min). f, Sample after extended incubation of the four-enzyme reaction (t = 360 min). The total-ion chromatograms (TICs) recorded in the negative-ion mode from the MS-MS fragmentation of the quasi-molecular ions [M-H]- of SQ and SF and SFP, DHAP, SLA and DHPS, from a representative experiment (n = 5) are shown.

For representative MS-MS fragmentation patterns of the [M-H]- ions of SQ and SF, SFP, and SLA and DHPS, see the Extended Data Figures 5, 6 and 7, respectively.

Together, the results show that the SQ-pathway in E. coli K-12 (Figure 2a) does not involve a desulfonation reaction and that no substrate-level phosphorylation of the sulfonated C3

intermediate occurs, which has been used previously (Roy et al. 2003) as a default hypothesis.

Furthermore, we deduce that there are ten genes in the gene cluster (Figure 2b). The core pathway comprises a SQ transporter (for example, b3876, YihO), SQ isomerase (b3880, YihS), SF kinase (b3883, YihV), SFP aldolase (b3881, YihT), SLA reductase (b3882, YihU) and a DHPS exporter (for example, b3877, YihP), which could be under the putative control of repressor b3884 (YihW). We propose a sulfolipid porin (b3875, OmpL), a sulfolipid

(32)

glucosidase (b3878, YihQ), and an epimerase (b3879, YihR) to funnel other SQ derivatives into the pathway, for example, the whole sulfolipid (see Extended Data Figure 8).

The gene cluster is found in at least 1,009 (>91%) of the 1,110 genome sequences of commensal E. coli, as well as pathogenic E. coli (for example, EHEC) strains, that were available in November 2013 (finished and draft genome sequences) in the Integrated Microbial Genomes (IMG) and Human Microbiome Project (HMP) databases (that is, gene clusters with syntenic yihTUVW and collinear homologs of yihSRQPO and ompL in variable order). Hence, the gene cluster is a feature of the core-genome of E. coli species. It can also be found in a wide range of other Enterobacteriaceae (for example, Chronobacter sakazakii ATCC BAA-894, Klebsiella oxytoca 10-5242, Pantoea ananatis LMG 20103 and Salmonella enterica LT2). We therefore suspect that the pathway has a significant role in bacteria in the alimentary tract of all omnivores and herbivores, that the pathway occurs in excrement from these animals, and in plant pathogens, which would explain part of the widespread occurrence of microbial degradation observed (Martelli 1967, Roy et al. 2003, Denger et al. 2012).

SQ is produced in huge amounts in nature and, thus, represents a significant proportion of the organic sulfur cycle (Harwood and Nicholls 1979), and it is degraded in similar amounts by both bacteria (Martelli 1967, Roy et al. 2003, Denger et al. 2012) and algae (Sugimoto et al.

2007), or it would accumulate in the environment. We see here that the Enterobacteriaceae use one pathway (Figure 2a) to initiate degradation of SQ, and that a community is required for complete degradation (Figure 1b) (Denger et al. 2012). This covers a variety of habitats, but we know that other pathways exist. A previous paper (Roy et al. 2003) presented evidence for SQ dehydrogenase, which we also observe in our SQ-using strain of Pseudomonas putida (A.-K.F. unpublished observations). Notably, another group (Martelli 1967) reported complete SQ degradation, including desulfonation, in a single organism; however, this organism has been lost (Cook and Denger 2002).

In summary, we have established that sulfoglycolysis, which was named but not defined in a previous report (Benson and Shibuya 1961), converts SQ to DHPS in the most widely-studied prokaryotic model organism, E. coli K-12, representing many Enterobacteriaceae (Figure 2a).

We have identified a gene cluster in E. coli K-12 (Figure 2b), which encodes the pathway.

The core pathway, for SQ, involves four newly discovered enzymes, two newly identified transporters and three newly characterized intermediates (Figures 2a, b). We know that the pathway is regulated (Extended Data Figures 1 and 2) and we suspect that it includes the degradation of the whole sulfolipid (Extended Data Figure 8). The pathway represents a substantial part of the biogeochemical sulfur cycle, and the pathway is likely to have a

(33)

significant role in bacteria in the alimentary tract of all omnivores and herbivores, and in plant pathogens. We and others (Martelli 1967, Roy et al. 2003, Denger et al. 2012) anticipate other degradative pathways for SQ in nature; for example, in bacteria of all marine, freshwater and terrestrial habitats where SQ is produced and degraded. We now provide the tools to elucidate these degradative pathways.

METHODS SUMMARY

SQ and DHPS were synthesized chemically and identified by NMR and mass spectrometry (Mayer et al. 2010, Denger et al. 2012). Cultivation, preparation of cell-free extracts, enzyme purification, 2D-PAGE and PF-MS, RNA preparation and RT-PCR, and expression and purification of His-tagged proteins, are described in the Methods. SQ, SF, SFP, SLA, DHAP and DHPS were separated using hydrophilic interaction liquid chromatography (HILIC) (Denger et al. 2012) and detected by an evaporative light scattering detector (ELSD) (Denger et al. 2012) or electrospray ionization (ESI)-time-of-flight (TOF)-MS or ESI-iontrap-MS (see Methods). The enzyme reaction mixture (Figure 3) was 3 mM SQ in 50 mM ammonium acetate buffer (pH 8.7), and 8 mM ATP, 0.5 mM MgCl2 and 8 mM NADH supplemented with the corresponding enzymes (each 50 µg ml-1).

ACKNOWLEDGEMENTS

We thank E. Deuerling for substrain MG1655, J. Klebensberger for substrain BW25113 and its knock-out mutants, and K. Leitner for help with growth experiments. The work of M.W.

and A.-K.F. was supported by the Konstanz Research School Chemical Biology (KoRS-CB), the work of C.M. by German Research Foundation (DFG) grants (MA2436/4 and SFB766/A15) and by the Baden-Württemberg Stiftung (P-BWS-Glyko11), and the work of D.Sc. by a DFG grant (SCHL 1936/1-1) and by the University of Konstanz and the Konstanz Young Scholar Fund.

(34)

METHODS

Chemicals

SQ and DHPS were synthesized chemically and identified by NMR and MS as described previously (Mayer et al. 2010, Denger et al. 2012). Dihydroxyacetone phosphate dilithium salt, D-fructose 6-phosphate disodium salt hydrate, D-glucose 6-phosphate disodium salt hydrate, and D-fructose 1,6-bisphosphate trisodium salt octahydrate were from Sigma. Other biochemicals (NADH, NADPH, NAD+, NADP+, ATP, and ADP) were purchased from Sigma, Fluka, Merck or Biomol.

Bacteria and growth conditions

Escherichia coli K-12 substrains W3100 (DSM 5911, ATCC 27325) and DH1 (DSM 4235, ATCC 33849), and Cupriavidus pinatubonensis JMP134 (DSM 4058) (Sato et al. 2006) were purchased from the Leibniz Institute DSMZ - Deutsche Sammlung von Mikroorganismen und Zellkulturen GmbH. E. coli K-12 substrain MG1655 (DSM

18039)

was a gift from E. Deuerling, and E. coli K-12 substrain BW25113 and its single-gene knockouts from the E. coli Keio Knockout Collection (Baba et al. 2006) were a gift from J. Klebensberger. The growth medium was a phosphate-buffered mineral salts medium (Thurnheer et al. 1986) (pH 7.2) with SQ or glucose as the sole carbon sources. Cultures were inoculated (1%) with pre- culture grown with the same substrate, and grown aerobically at 30°C. Cultures in 3 ml volume were grown in screw-cap tubes (30 ml) in a roller, and cultures in the 50 ml or 200 ml volume in capped Erlenmeyer flasks (0.3 or 1.0 liter volume, respectively) on a horizontal shaker; for the latter, 0.8-ml samples were taken at intervals to determine optical density (attenuance D at 580 nm; D580nm) and substrate and product concentrations (HILIC-HPLC, see below). For the growth experiments to demonstrate mineralization of SQ (see Figure 1), E.

coli K-12 substrain MG1655 was grown with SQ (4 mM; 50-ml scale), and after growth had been completed, the cellular biomass was removed from the culture fluid by centrifugation (20,000g, 30 min, 4°C) followed by filter-sterilization (pore size, 0.2 µm). The culture fluid was then inoculated with C. pinatubonensis JMP134. During the growth experiments, samples were taken at intervals to monitor the growth (D580) and to determine total protein, substrate, and product concentrations (see below). All growth experiments were replicated twice (n = 3).

(35)

Preparation of crude extract, soluble fraction and enzyme enrichment

E. coli cells from growth experiments with SQ or glucose were harvested at an D580 of approximately 0.4 by centrifugation (20,000g, 15 min, 4°C) and disrupted by three passages through a chilled French pressure cell (140 megapascals (MPa); Aminco) in the presence of DNase (25 µg ml-1). Cell debris was removed by centrifugation (11,000g, 5 min, 4°C) and the membrane fragments collected by ultracentrifugation (70,000g, 1 h, 4°C); the supernatant was called the soluble fraction. For enzyme enrichment, soluble fraction was loaded onto an anion-exchange chromatography column (MonoQ HR 10/10 column, Pharmacia) equilibrated with 20 mM Tris/H2SO4 buffer, pH 9.0, at a flow rate of 1 ml min-1, and bound proteins eluted by a linear Na2SO4 gradient (from 0 M to 0.2 M in 45 min, and to 0.5 M in 10 min) and fractions (2 ml) collected; the SLA reductase activity eluted at about 0.12 M Na2SO4.

Two dimensional gel electrophoresis and peptide fingerprinting-mass spectrometry

2D-PAGE and PF-MS were done according to our previously published protocols (Schmidt et al. 2013). In brief, soluble protein fractions from E. coli cells grown with SQ or glucose (see above) were desalted (PD-10 Desalting Columns, GE Healthcare Life Sciences) and precipitated by addition of acetone (four volumes 100% acetone, -20°C, overnight); each 1 mg of precipitated protein was solubilized in rehydration buffer (300 µl) and loaded onto isoelectric focusing (IEF) strips (BioRad ReadyStrip IPG system) overnight; IEF involved a voltage ramp to 10,000 V during 3 h, and a total focusing of 40,000 Volt-hours (Vh); the strips were equilibrated in SDS-equilibration buffers I and II (with DTT and iodoacetamide, respectively) and placed onto SDS-PAGE gels using an overlay of SDS-gel buffer solidified with agarose (0.5%); SDS-PAGE gels contained 12% polyacrylamide (no stacking gel), and were stained with Coomassie brilliant blue R-250 (Laemmli 1970). Stained protein spots of interest were excised from gels and submitted to PF-MS at the Proteomics Facility of the University of Konstanz to identify the corresponding genes; the MASCOT engine (Matrix Science, London, UK) was used to match each peptide fingerprint against a local database of all predicted protein sequences of the annotated E. coli K-12 substrain MG1655 genome (IMG version 2011-08-16).

Total RNA preparation and PCR with reverse transcription

RNA preparation and RT-PCR were done according to our previously published protocols (Weiss et al. 2012). In brief, cells were grown in the appropriate selective medium (3 ml) and

(36)

harvested in the mid-exponential growth phase (D580 ≈ 0.3); the cell pellets were stored at -20°C in RNAlater RNA stabilization solution (Applied Biosystems); total RNA was prepared using the E.Z.N.A. Bacterial RNA Kit (Omega Bio-Tek) following the manufacturer’s instructions; the RNA preparation (40 µl) was treated with RNase-free DNase (2 U, 30 min, 37°C) (Fermentas). For complementary DNA (cDNA) synthesis, the Maxima reverse transcriptase (Fermentas) was used following the manufacturer’s instructions; the reactions contained 0.2 µg total RNA and 20 pmol sequence-specific primer (see below). PCR reactions (20 µl volume) were done using Taq DNA polymerase (Fermentas) and the manufacturer’s standard reaction mixture (including 2.5 mM MgCl2); cDNA from reverse transcription reactions was used as template (2 µl of reverse transcription reaction mixture), or genomic DNA (4 ng DNA) for PCR-positive controls, or non-reverse transcribed total RNA (2 µl) for the confirmation of an absence of DNA impurities in the RNA preparations.

Heterologous expression and purification of His-tagged proteins

Heterologous expression of candidate genes and purification of the recombinant proteins were done according to our previously published protocol (Felux et al. 2013). In brief, chromosomal DNA was isolated using the Illustra bacteria genomicPrep Mini Spin Kit (GE Healthcare) and the target genes amplified by PCR using Phusion HF DNA Polymerase (Finnzymes) and the primer pairs given below; the PCR conditions were 30 cycles of 18 s denaturation at 98°C, 20 s annealing at 58°C, and 60 s elongation at 72°C for gene b3880, or 45 s elongation at 72°C for genes b3881, b3882, and b3883; the PCR products were then separated by agarose-gel electrophoresis, excised, and purified using the QIAquick gel extraction kit (Qiagen), and ligated into the amino-terminal His6-tag expression vector pET100 (Invitrogen); correct integration of the inserts was confirmed by sequencing (GATC- Biotech). For expression, BL21 Star (DE3) OneShot E. coli cells (Invitrogen) were transformed with the constructs and grown at 37°C in lysogeny broth medium containing 100 mg l-1 ampicillin; at an D580 ≈ 0.6, the cultures were induced by addition of 0.5 mM IPTG (isopropyl-β-D-thiogalactoside), and the cells grown for additional 4 to 5 h at 20°C, collected by centrifugation (15,000g, 20 min, 4°C), and stored frozen (-20°C). Cells were resuspended in buffer A (20 mM Tris/HCl, pH 8.0, 100 mM KCl) that contained 50 µg ml-1 DNase I, and disrupted by four passages through a pre-cooled French pressure cell (140 MPa). The cell extracts were centrifuged (15,000g, 10 min, 4°C) and ultra-centrifuged (70,000g, 1 h, 4°C), and the soluble protein fractions loaded onto 1-ml Ni2+-chelating Agarose affinity columns (Macherey-Nagel) pre-equilibrated with buffer A (see above). After a washing step (30 mM

(37)

imidazole in buffer A), the His-tagged proteins were eluted (200 mM imidazole in buffer A), concentrated in a Vivaspin concentrator (Sartorius), and, after addition of 30% glycerol (v/v), stored in aliquots at -20°C.

Enzyme assays

SLA reductase activity was assayed photometrically at 365 nm in 1-ml cuvettes in 50 mM Tris/HCl buffer, pH 9.0, with 1 mM NAD+ and 5 mM DHPS as substrates. The reaction was started with the addition of protein (0.01 – 0.1 mg ml-1) and the reduction of NAD+ was recorded. Enzyme assays with recombinant proteins (each 50 µg protein ml-1) for analysis by HILIC-HPLC were carried out in the 1 ml volume in 50 mM ammonium acetate buffer, pH 8.7, stirred at room temperature (approximately 20 – 23°C); samples were taken at intervals, for which the reactions were stopped by addition of 30% acetonitrile. SQ (3 mM) and recombinant isomerase were incubated for 60 min, after which ATP (8 mM), MgCl2

(0.5 mM), and recombinant kinase were added. After additional 60 min, recombinant aldolase was added, and after another 60 min NADH (8 mM) and the recombinant reductase.

Analytical methods

Total protein was determined according to a protocol based on the method reported previously (Kennedy and Fewson 1968), and soluble protein by protein dye binding (Bradford 1976), each using bovine serum albumin (BSA) as the standard. Sulfate release during growth was quantified turbidimetrically (Sörbo 1987) as a suspension of BaSO4. For HPLC-ESI-MS-MS, an Agilent 1100 HPLC system fitted with a ZIC-HILIC column (5 µm, 200 Å, 150 x 4.6 mm;

Merck) was connected to an LCQ ion trap mass spectrometer (Thermofisher). The HPLC conditions for the LCQ ion trap were: From 90% B to 65% B in 25 min, 65% B for 10 min, in 0.5 min back to 90% B, 90% B column equilibration for 9.5 min; solvent A, 90% 0.1 M NH4Ac, 10% acetonitrile; solvent B, acetonitrile; flow rate, 0.75 ml min-1. The mass spectrometer was run in ESI negative mode. The retention times and ESI-MS-MS fragmentation patterns of the analytes were observed as follows: SQ retention time, 25.4 min;

SQ ESI-MS m/z (per cent base-peak) 243 (100); SQ ESI-MS-MS of [M-H]- 243: 243 (4), 225 (11), 207 (34), 183 (100), 153 (54), 143 (1), 123 (16), 101 (8), 81 (6). SF retention time, 21.9 min; SF ESI-MS, 243 (100); SF ESI-MS-MS of [M-H]- 243: 243(1), 225 (37), 207 (38), 183 (21), 153 (100), 143 (3), 123 (24), 101 (13), 81 (5). SFP retention time, 33.4 min; SFP ESI- MS, 323 (100); SFP ESI-MS-MS of [M-H]- 323: 305 (34), 287 (3), 233 (2), 225 (100), 207 (32), 153 (4). SLA retention time, 21.0 min; SLA ESI-MS, 153 (100); SLA ESI-MS-MS of

(38)

[M-H]- 153: 153 (9), 81 (100), 71 (18). DHPS retention time, 18.0 min; DHPS ESI-MS, 155 (100), 95 (4); DHPS ESI-MS-MS of [M-H]- 155: 155 (100), 137 (18), 95 (40). DHAP retention time, 30.2 min; DHAP ESI-MS: 169 (100); DHAP ESI-MS-MS of [M-H]- 169: 169 (2), 125 (2), 97 (100). HPLC for ESI-TOF-MS (MicrO-TOF II, Bruker) involved the same column and gradient system, but a different gradient program (from 90% B to 65% B in 20 min, and further to 55% B in 20 min), which resulted in retention times (see Extended Data Figure 4) of 22.4 min for SQ, 21.8 min for SF, 16.3 min for fructose, 18.5 min for glucose, 35.2 min for fructose-6-phosphate, and 39.8 min for glucose-6-phosphate.

PCR primers

Primers were purchased from Microsynth (Balgach). The sequences of the primers pairs (for/rev) for RT-PCR (see above) were (product length in bp): b3879 forward, 5’-CCTTATGGCGTGGGTATTCATCC-3’, b3879 reverse, 5’-TTAGGCGGGCAACTCAT- AGGTTC-3’ (353); b3880 forward, 5’-ACGCGGTGGAAGCTTTCTTGAT-3’, b3880 reverse, 5’-CACGGTGGCGTTAAACAGACCTT-3’ (332); b3881 forward, 5’-TGTCGCC- GCCGATGAGTTC-3’, b3881 reverse, 5’-CTTTGTAGAGGTCAGCGCCACTGT-3’ (320);

b3882 forward, 5’-GGCGCAGGCCGCTAAAGA-3’, b3882 reverse, 5’-AAGATTCAGG- GCTTCGCACAAAA-3’ (439); b3883 forward, 5’-GGCACGACGGCGCTAAAAA-3’, b3883 reverse, 5’-TGACTCCGCTAAATCCCCACTTG-3’ (374); b0720 forward, 5’-CGCT- GGCGGCGTTCTATCA-3’, b0720 reverse, 5’-ATTTTCAGCGCCGCTTCGTTAG-3’ (403).

The sequences of the primer pairs for TOPO-cloning and heterologous expression (see above) were (the directional overhang is underlined): b3880 forward, 5’-CACCGGAATGAA- ATGGTTTAACACCCTAAG-3’, b3880 reverse, 5’-AACCCGCACC-CTATTTTCAG-3’;

b3881 forward, 5’-CACCATGAATAAGTACACCATCAACGACATT-ACG-3’, b3881 reverse, 5’-ACCATTTCATTCCTTTTATCCTCATCTT-3’; b3882 forward, 5’-CACCATG- GCAGCAATCGCGTTTATCG-3’, b3882 reverse, 5’-CGCGTAATGTCGTTGATGG- TGTA-3’; b3883 forward, 5’-CACCATGATTCGTGTTGCTTGTGTAGGT-3’, b3883 reverse, 5’-TGAAAATTCCTCGAAAAACCATCA-3’.

Genome analyses

Analysis of genomes for orthologous gene clusters was carried out through the gene cassette search and neighborhood regions search options of the Integrated Microbial Genomes (IMG) and IMG Human Microbiome Project (IMG HMP) platforms (http://img.jgi.doe.gov/). Basic

(39)

sequence analyses were done using NCBI’s BLAST tools (http://blast.ncbi.nlm.nih.gov) and the Lasergene DNAstar software package (www.dnastar.com).

Enzyme nomenclature

We suggest that sulfolactaldehyde reductase belongs to the NC-IUBMB subgroup EC 1.1.1., with the name sulfolactaldehyde 3-reductase (systematic name 2,3-dihydroxypropane-1- sulfonate:NAD+ 3-oxidoreductase). The sulfofructose kinase would then belong to EC 2.7.1., with the name sulfofructose 1-kinase (systematic name ATP:6-deoxy-6-sulfofructose 1- phosphotransferase). The sulfofructosephosphate aldolase would belong to EC 4.1.2., with the name sulfofructosephosphate aldolase (systematic name 6-deoxy-6-sulfofructose 1-phosphate 2-hydroxy-3-oxopropane-1-sulfonate-lyase (glycerone phosphate forming)). Sulfoquinovose isomerase would belong to EC 5.3.1., with the name sulfoquinovose isomerase (systematic name 6-deoxy-6-sulfoglucose aldose-ketose-isomerase).

(40)

EXTENDEND DATA FIGURES AND TABLES

Extended Data Figure 1. Soluble proteins in glucose- or SQ-grown cells of E. coli K-12 MG1655 separated by 2D-PAGE. All prominent protein spots on the gel from SQ-grown cells that suggested inducibly expressed proteins were excised and submitted to PF-MS (see Extended Data Table 1). The PF-MS identifications were replicated in an independent growth experiment and gel electrophoresis run.

Extended Data Table 1. Identifications by peptide fingerprinting-mass spectrometry of protein spots excised from 2D-PAGE gels of SQ-grown E. coli cells

Protein spots were sorted according to their apparent molecular mass on the gel (see Extended Data Figure 1).

Referenzen

ÄHNLICHE DOKUMENTE

Lanes 1 to 6 of the gel contain the amplification products of each of the three separate PCRs: specific 0.97-kb fragment of K-12 derivatives, resulting from amplification with

Upon activation, autoproteolytic cleavage of caspase-1 in the inammasome complex induces a lytic form of programmed cell death called pyroptosis [139, 252].. Pyroptosis

As our previous work showed that wild-type ribosomes were stable during constant growth in turbidostat and there was no growth in the stationary phase, we decided to measure

The toxicity of cholate and its degradation intermediates with a steroid structure indicates that strain Chol1 requires a strategy to minimize these toxic effects during growth

Sulfoquinovose degraded by two pure cultures to 3-sulfolactate or DHPS, and the degradation of these two compounds by two SQ- negative pure cultures to yield

As the pathway for bile salt degradation merges with metabolic pathways for bacterial transformation of other steroids, such as testosterone and cholesterol, this

Thus, propionyl-CoA and the presumptive intermediates of the propionyl-CoA oxidation pathway(s) may also exert an effect in cholate-grown cells of strain

Dithionite-reduced benzyl or methyl viologen have previously been used as low-potential reductants to provide exergonic redox conditions for driving the reverse reaction