28.12.2013 Views

genomewide characterization of host-pathogen interactions by ...

genomewide characterization of host-pathogen interactions by ...

genomewide characterization of host-pathogen interactions by ...

SHOW MORE
SHOW LESS

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

G E N O M E W I D E C H A R A C T E R I Z A T I O N<br />

O F H O S T - P A T H O G E N I N T E R A C T I O N S<br />

B Y T R A N S C R I P T O M I C A P P R O A C H E S<br />

I N A U G U R A L D I S S E R T A T I O N<br />

ZUR<br />

ERLANGUNG DES AKADEMISCHEN GRADES<br />

DOCTOR RERUM NATURALIUM (DR. RER. NAT.)<br />

AN DER MATHEMATISCH-NATURWISSENSCHAFTLICHEN FAKULTÄT<br />

DER ERNST-MORITZ-ARNDT-UNIVERSITÄT GREIFSWALD<br />

VORGELEGT VON MAREN DEPKE<br />

GEBOREN AM 11. 01. 1978<br />

IN HAMBURG<br />

GREIFSWALD, DEN 24. 09. 2010


Dekan:<br />

Pr<strong>of</strong>. Dr. rer. nat. Klaus Fesser<br />

1. Gutachter : Pr<strong>of</strong>. Dr. rer. nat. Uwe Völker<br />

2. Gutachter: Pr<strong>of</strong>. Dr. rer. nat. Christiane Wolz<br />

Tag der Promotion: 22. 12. 2010


Maren Depke<br />

C O N T E N T S<br />

GENOMEWIDE CHARACTERIZATION OF HOST-PATHOGEN INTERACTIONS<br />

BY TRANSCRIPTOMIC APPROACHES 1<br />

CONTENTS 3<br />

ZUSAMMENFASSUNG DER DISSERTATION 5<br />

SUMMARY OF DISSERTATION 9<br />

INTRODUCTION 13<br />

INFECTIOUS DISEASES AND IMMUNE SYSTEM 13<br />

Aspects <strong>of</strong> Innate Immunity 13<br />

Aspects <strong>of</strong> Adaptive Immunity 16<br />

Modulation <strong>of</strong> Immune Reactions 18<br />

STAPHYLOCOCCUS AUREUS 20<br />

General Features <strong>of</strong> S. aureus 20<br />

S. aureus, a Commensal and Opportunistic Pathogen 21<br />

S. aureus Virulence Factors 22<br />

Mechanisms <strong>of</strong> S. aureus Adaptation to its Environment 28<br />

Increasing Importance and Danger <strong>of</strong> S. aureus Infections 29<br />

STUDIES OF HOST-PATHOGEN INTERACTIONS 31<br />

Model Systems for Studies <strong>of</strong> Host Reactions Potentially Influencing the Outcome <strong>of</strong> Infections 31<br />

Model Systems for Studies <strong>of</strong> Host-Pathogen Interactions 35<br />

Questions and Aims <strong>of</strong> the Studies Described in this Thesis 37<br />

MATERIAL AND METHODS 39<br />

LIVER GENE EXPRESSION PATTERN IN A MOUSE PSYCHOLOGICAL STRESS MODEL 39<br />

KIDNEY GENE EXPRESSION PATTERN IN AN IN VIVO INFECTION MODEL 43<br />

GENE EXPRESSION PATTERN OF BONE-MARROW DERIVED MACROPHAGES AFTER INTERFERON-GAMMA TREATMENT 47<br />

HOST CELL GENE EXPRESSION PATTERN IN AN IN VITRO INFECTION MODEL 51<br />

PATHOGEN GENE EXPRESSION PROFILING 55<br />

Growth Media Comparison Study 55<br />

In vitro Infection Experiment Study 57<br />

Tiling Array Expression Pr<strong>of</strong>iling 60<br />

3


Maren Depke<br />

Contents<br />

RESULTS 63<br />

LIVER GENE EXPRESSION PATTERN IN A MOUSE PSYCHOLOGICAL STRESS MODEL 63<br />

KIDNEY GENE EXPRESSION PATTERN IN AN IN VIVO INFECTION MODEL 75<br />

GENE EXPRESSION PATTERN OF BONE-MARROW DERIVED MACROPHAGES AFTER INTERFERON-GAMMA TREATMENT 91<br />

HOST CELL GENE EXPRESSION PATTERN IN AN IN VITRO INFECTION MODEL 111<br />

PATHOGEN GENE EXPRESSION PROFILING 131<br />

Growth Media Comparison Study 131<br />

In vitro Infection Experiment Study 136<br />

DISCUSSION AND CONCLUSIONS 171<br />

LIVER GENE EXPRESSION PATTERN IN A MOUSE PSYCHOLOGICAL STRESS MODEL 171<br />

KIDNEY GENE EXPRESSION PATTERN IN AN IN VIVO INFECTION MODEL 178<br />

GENE EXPRESSION PATTERN OF BONE-MARROW DERIVED MACROPHAGES AFTER INTERFERON-GAMMA TREATMENT 182<br />

HOST CELL GENE EXPRESSION PATTERN IN AN IN VITRO INFECTION MODEL 190<br />

PATHOGEN GENE EXPRESSION PROFILING 197<br />

REFERENCES 207<br />

PUBLICATIONS 233<br />

AFFIDAVIT / ERKLÄRUNG 235<br />

ACKNOWLEDGMENTS 237<br />

4


Maren Depke<br />

Z U S A M M E N F A S S U N G D E R D I S S E R T A T I O N<br />

Mensch und Tier sind regelmäßig Mikroorganismen ausgesetzt. In der Interaktion mit <strong>pathogen</strong>en<br />

Mikroorganismen entwickelten sich Abwehrmechanismen, die entweder Infektionen verhindern<br />

oder sie zu überwinden helfen. Parallel dazu erwarben Pathogene Mechanismen, um der<br />

Abwehr ihres Wirts zu entgehen. Außer der wirtseigenen Immunregulation und den Faktoren der<br />

Pathogene üben weitere Faktoren wie körperliche Anstrengung und die psychische Verfassung<br />

des Wirts Einfluß auf das Immunsystem aus. Während kurze Streßepisoden sogar die Immunantwort<br />

fördern können, kann sie durch zu lang andauernde Streßphasen negativ beeinflußt<br />

werden. Doch nicht nur die Immunantwort, sondern auch metabolische Prozesse unterliegen<br />

Modifikationen durch solche Stressoren. Deshalb können Untersuchungen zu Wirt-Erreger-<br />

Wechselwirkungen helfen, Mechanismen aufzuklären, die entweder für Wirt oder Pathogen von<br />

Vorteil sind, und dazu beitragen, Interventionsstrategien im Fall von Infektionskrankheiten zu<br />

etablieren. Diese Dissertation beschreibt die Ergebnisse aus Transkriptomstudien zu verschiedenen<br />

Aspekten von Wirt-Erreger-Wechselwirkungen.<br />

Zunächst wurde das Lebergenexpressionspr<strong>of</strong>il aus einem Mausmodell für chronischen,<br />

psychologischen Streß verwendet, um den Einfluß von Streß auf Metabolismus und Immunantwort<br />

der Tiere zu verdeutlichen. Psychische und physische Stressoren können neuroendokrine,<br />

immunologische, verhaltensbezogene und metabolische Funktionen stören. Vor kurzem wurde<br />

von Kiank et al. publiziert, daß BALB/c-Mäuse ein schwere systemische Immunsuppression,<br />

neuroendokrine Störungen und depressionsähnliches Verhalten entwickeln, wenn sie in einem<br />

Modell für starken, chronischen, psychischen Streß über 4,5 Tage periodisch akustischem Streß in<br />

Kombination mit Bewegungseinschränkung ausgesetzt wurden (Kiank et al. 2006; Brain Behav<br />

Immun. 20(4):359). Außerdem litten diese Mäuse unter deutlichem Gewichtsverlust. Um Gründe<br />

dafür aufzuklären, wurde das Genexpressionspr<strong>of</strong>il der Leber, die eine Hauptrolle im St<strong>of</strong>fwechsel<br />

übernimmt, analysiert. Die Leber übt außerdem eine Wächterfunktion zwischen dem<br />

Verdauungstrakt und dem Blutsystem aus. Deshalb wurde in dem Genexpressionsdatensatz<br />

zusätzlich der Einfluß von psychischem Streß auf immunregulierende Prozesse untersucht.<br />

Bereits nach einer einzelnen, akuten Streßphase wies das hepatische Genexpressionsmuster<br />

deutliche Veränderungen auf. Auch wenn zu diesem Zeitpunkt noch keine metabolischen Veränderungen<br />

sichtbar wurden, begann dennoch eine Genexpressionskaskade, die zu den beobachteten<br />

Störungen führte, nachdem der Streß die chronische Phase erreicht hatte. Dort waren<br />

dann besonders st<strong>of</strong>fwechselbezogene Gene in ihrer Expression verändert. Die differentielle<br />

Expression betraf Kohlenhydrat-, Aminosäure- und Fettmetabolismus. Es wurde gezeigt, daß<br />

chronischer Streß in weiblichen BALB/c-Mäusen zu einem hypermetabolischen Syndrom einschließlich<br />

Auslösung von Gluconeogenese und Hypercholesterinämie und dem Verlust von<br />

essentiellen Aminosäuren führte. Des weiteren deckte diese Analyse eine veränderte Expression<br />

von Genen der Immunantwort auf. Darin war die Auslösung einer Akute-Phase-Antwort, aber<br />

auch von immunsupprimierenden Abläufen und die Unterdrückung von hepatischer Antigenpräsentation<br />

enthalten. In chronisch gestreßten Mäusen wurde gesteigerte Leukozyteneinwanderung,<br />

verstärkter oxidativer Streß, der aber auch mit gegenregulatorischen Expressionsveränderungen<br />

einherging, sowie Apoptose detektiert.<br />

5


Maren Depke<br />

Zusammenfassung der Dissertation<br />

Die Experimente in der Studie am Modell für psychischen Streß wurden noch ohne den zusätzlichen<br />

Einfluß eines <strong>pathogen</strong>en Erregers durchgeführt, der aber in der zweiten Studie berücksichtigt<br />

wurde. Hierbei wurde der Einfluß einer intravenösen Staphylokokkeninfektion auf die<br />

Nierengenexpression des Wirts in einem weiteren in vivo Mausmodell analysiert, wobei der<br />

Wildtypstamm Staphylococcus aureus RN1HG und seine isogene sigB-Mutante eingesetzt<br />

wurden. S. aureus, ein Gram-positives Bakterium, besiedelt als persistierender Kommensale den<br />

vorderen Nasenraum von ca. 20 % der menschlichen Bevölkerung. Normalerweise bewirkt die<br />

Besiedlung mit S. aureus keine Erkrankung. Andererseits kann S. aureus aber auch für ein breites<br />

Spektrum an Erkrankungen verantwortlich sein, das von schwachen bis schweren lokalen Infektionen<br />

der Haut oder Rachenschleimhaut über Infektionen innerer Organe (z. B. Endokarditis,<br />

Osteomyelitis) bis zu systemischen Erkrankungen wie Sepsis geht. In das Blut kann S. aureus nach<br />

Verletzungen oder durch medizinische Hilfsmittel wie Katheter gelangen. Ein einfaches Modell,<br />

um solche Blutsysteminfektionen zu imitieren, ist die i. v.-Infektion von Mäusen. Die Wirtsreaktion<br />

kann dabei mit physiologischen, immunologischen und molekularen Messungen aufgezeichnet<br />

werden. In dieser Studie wurde die Transkriptomanalyse auf Mausnierenproben angewandt.<br />

Obwohl Literaturangaben <strong>of</strong>t von ähnlicher Virulenz von sigB-defizienten Mutanten und Wildtypstämmen<br />

berichten, könnte sich der Mechanismus der Pathogenese – zum Teil auch in<br />

Abhängigkeit von dem gewählten Infektionsmodell – zwischen beiden unterscheiden. Deshalb<br />

sollte mit dieser Studie untersucht werden, ob die Deletion von sigB zu einer veränderten Wirtsantwort<br />

während der Infektion führt. Das Genexpressionspr<strong>of</strong>il in infiziertem Nierengewebe war<br />

sehr gut reproduzierbar. Der Vergleich mit nichtinfizierten Kontrollen zeigte eine starke, proinflammatorische<br />

Reaktion der Niere, die z. B. Signalweitergabe sogenannter „Toll-like“-Rezeptoren,<br />

Komplementsystem, Antigenpräsentation, Interferon- und IL-6-, aber auch gegenregulatorische<br />

IL-10-Signalwege einschloß. Die Studie konnte keine Unterschiede im Mechanismus der<br />

Pathogenese der S. aureus-Stämme RN1HG und seiner isogenen sigB-Mutante belegen, da sich<br />

die Wirtsantwort in beiden Fällen nicht unterschied. Falls solche Unterschiede tatsächlich<br />

existieren sollten, sind sie womöglich transienter Natur und nur zu früheren Zeitpunkten auffällig.<br />

Effekte von SigB könnten in der in vivo Infektion auch von dem verflochtenen Regulationsmuster<br />

anderer Regulatoren überlagert sein. Des weiteren besteht die Möglichkeit, daß SigB in vivo gar<br />

nicht aktiv ist, wodurch die ähnliche Wirtsreaktion auf Wildtyp und Mutante erklärbar wäre.<br />

Möglicherweise besitzt SigB weniger Eigenschaften eines Virulenzfaktors als vielmehr eines<br />

Virulenzmodulators, der in vivo die Feinabstimmung der bakteriellen Reaktionen übernimmt,<br />

oder SigB ist in speziellen Nischen während einer Infektion von Bedeutung. Solch eine Funktion<br />

würde das Fehlen nachweisbarer Unterschiede zwischen der Wirtsreaktion auf S. aureus RN1HG<br />

und seine isogene sigB-Mutante erklären.<br />

Expressionsstudien an Gewebeproben aus in vivo Modellen zeichnen direkt den relevanten<br />

physiologischen Zustand im Zusammenhang mit allen komplexen Interaktionen und Einflüssen<br />

auf und liegen medizinischen Fragestellungen am nächsten. Dennoch ist es schwer, die einzelnen<br />

Anteile zu unterscheiden, da es sich bei Gewebe immer um eine Mischung verschiedener Zelltypen<br />

handelt, die sogar gegensätzliche Reaktionen aufweisen können. Deshalb wurden zusätzlich<br />

in vitro Modelle untersucht, die sich auf einen einzelnen, definierten Zelltyp fokussierten.<br />

Makrophagen sind in die erste Reaktion auf eine Infektion einbezogen. Sie nehmen, zusammen<br />

mit dendritischen Zellen, eine zentrale Position im angeborenen Immunsystem ein. Durch<br />

ihre Funktion als Wächter und Phagozyten sind sie maßgeblich an der Beseitigung von Infektionen<br />

beteiligt. Peptide phagozytierter Antigene werden durch sie auf MHC-II-Komplexen für<br />

6


Maren Depke<br />

Zusammenfassung der Dissertation<br />

Lymphozyten präsentiert, wodurch die Makrophagen auch an der Regulation der adaptiven<br />

Immunantwort teilhaben. Die Präparation von Knochenmarkstammzellen und ihre in vitro<br />

Differenzierung zu sogenannten “bone-marrow derived macrophages“ (BMM) stellt ein Modell<br />

zur Untersuchung der Reaktionen von Makrophagen dar, das immunologische Einflüsse, die vom<br />

Immunzustand des Tieres sogar unter standardisierten Laborbedingungen ausgehen können,<br />

umgeht. Bis vor kurzem wurden weitere unkontrollierbare Einflüsse durch die Verwendung von<br />

Serum, das undefinierte und variierende Substanzen enthält, mit dem Kulturmedium in die<br />

experimentelle Anordnung eingebracht. Um diese Ursache experimenteller Schwankungen zu<br />

vermeiden, wurde durch Eske et al. ein System der serumfreien Kultivierung von BMM eingeführt<br />

(Eske et al. 2009; J Immunol Methods. 342(1-2):13), das nun für die Untersuchung der Reaktion<br />

von BMM als drittem Teil dieser Dissertation angewendet wurde. Dabei wurden BMM verschiedener<br />

Mausstämme mit IFN-γ, einem Modulator der Makrophagenfunktion und einem der ersten<br />

Signale während der Initiation der Immunantwort, behandelt. Publizierte Experimente zeigten,<br />

daß BMM aus den Mausstämmen BALB/c oder C57BL/6 unterschiedlich auf die Konfrontation mit<br />

Burkholderia pseudomallei reagierten, insbesondere, wenn vorher eine Stimulation mit IFN-γ<br />

stattfand (Breitbach et al. 2006; Infect Immun. 74(11):6300). Auch weitere Studien wiesen Unterschiede<br />

zwischen den beiden Mausstämmen in vivo und in vitro nach. Vor dem Hintergrund der<br />

genetisch bedingten Unterschiede in der Reaktion der BMM wurden nun BALB/c- und C57BL/6-<br />

BMM mit IFN-γ stimuliert, um auf molekularer Ebene genomweit die Reaktion auf IFN-γ als<br />

Initialsignal zu bestimmen. Außerdem sollten in dieser Studie mögliche Unterschiede der<br />

Reaktion von BMM beider Stämme auf IFN-γ charakterisiert werden.<br />

Das Genexpressionspr<strong>of</strong>il zeigte nach der Behandlung mit IFN-γ in BMM beider Stämme hauptsächlich<br />

induzierte Genexpression. Darin wurden bekannte IFN-γ-Effekte wie die Induktion von<br />

Immunproteasom, Antigenpräsentation, Genen der Interferonsignalwege und von GTPasen/GTPbindenden<br />

Proteinen, sowie der induzierbaren Stickst<strong>of</strong>fmonoxidsynthase bestätigt. IFN-γabhängige<br />

Genexpressionsänderungen waren zwischen BALB/c- und C57BL/6-BMM in hohem<br />

Maße ähnlich. Sogar nur in BMM des einen Stamms signifikant verändert exprimierte Gene<br />

zeigten einen vergleichbaren Trend in BMM des anderen Stamms. Genexpressionsunterschiede<br />

zwischen BMM beider Mausstämme wurden sowohl in unbehandelten Kontroll-BMM als auch<br />

nach IFN-γ-Behandlung bestimmt. Dabei wiesen ca. 55 % bis 60 % der unterschiedlich stark<br />

exprimierten Gene eine höhere Expression in BALB/c-BMM auf, während die verbleibenden Gene<br />

stärker in C57BL/6-BMM exprimiert wurden. Vergleichbar zu der Beobachtung bei den IFN-γ-<br />

Effekten war auch bei den Unterschieden zwischen BMM der beiden Stämme ein ähnlicher Trend<br />

in beiden experimentellen Behandlungen, Kontrolle und IFN-γ-Stimulation, sichtbar. Die stammabhängig<br />

unterschiedlich exprimierten Gene schlossen immunrelevante und zelltodassoziierte<br />

Gene ein, aber die Abdeckung der funktionellen Gruppen war begrenzt. Phenotypische Unterschiede<br />

in der Reaktion von BALB/c- und C57BL/6-BMM wurden nach Literaturangaben zumeist<br />

in der Anwesenheit von IFN-γ und eines weiteren Stimulus wie LPS oder Infektion beobachtet.<br />

Um die molekularen Ursachen für die beobachteten Unterschiede in der Abtötung von<br />

Pathogenen oder der Produktion von Cytokinen zu bestimmen, müssen weitere Proben, die<br />

außer IFN-γ einem weiteren Stimulus ausgesetzt waren, eingeschlossen werden.<br />

Nicht nur Immunzellen bzw. Phagozyten kommen mit Pathogenen in Kontakt, sondern auch z. B.<br />

Epithel- oder Endothelzellen, die für den Erhalt von Struktur und Funktion von Wirtsorganen und<br />

-gewebe verantwortlich sind. Diese Zellen sind tatsächlich unter den ersten, die das Eindringen<br />

von Pathogenen erkennen und darauf reagieren. Auch wenn S. aureus den vorderen Nasen-<br />

7


Maren Depke<br />

Zusammenfassung der Dissertation<br />

bereich des Menschen besiedelt, kann das Bakterium Pneumonie verursachen, wenn es z. B.<br />

durch Einatmung oder medizinische Geräte in die Lunge gelangt. Die Bronchialepithelzellinie S9<br />

wurde als Modell für die in vitro Infektion mit Staphylokokken verwendet. Ein neues experimentelles<br />

System erlaubt die Untersuchung der Wirt-Erreger-Interaktion im Zusammenhang mit allen<br />

bakteriellen Faktoren, ob membrangebunden oder sezerniert, und vermeidet zusätzlich störende<br />

Effekte auf die Physiologie der Bakterien durch längeres Bearbeiten, Zentrifugation und Waschen<br />

(Schmidt et al. 2010; Proteomics. 10(15):2801). Das vierte Kapitel dieser Dissertation beschreibt<br />

S9-Genexpressionssignaturen nach in vitro Infektion mit S. aureus RN1HG für die zwei Zeitpunkte<br />

2,5 h und 6,5 h nach Beginn der Infektion. Zum frühen 2,5 h-Zeitpunkt wurde differentielle<br />

Expression nur für die kleine Anzahl von 40 Genen gemessen. Trotz der geringen Anzahl zeigten<br />

diese Gene eine beginnende pro-inflammatorische Antwort, z. B. durch die verstärkte Expression<br />

von Cytokinen (IL-6, IFN-β, LIF) oder Prostaglandinendoperoxidsynthase 2 (PTGS2), aber auch<br />

gegenregulatorische Prozesse wie die verstärkte Expression von CD274, Ligand für den immuninhibitorischen<br />

Rezeptor PD-1, waren erkennbar. Die Wirtsantwort verstärkte sich zum späteren<br />

6,5 h-Zeitpunkt dramatisch, an dem differentielle Expression für 1196 Gene detektiert wurde.<br />

Darin waren Cytokine, Signalwege der Rezeptoren für bakterielle Strukturen, Antigenpräsentation<br />

und an der Immunantwort beteiligte Gene (z. B. GBP, MX, APOL) enthalten. Negative<br />

Auswirkungen auf Wachstum und Proliferation traten noch stärker auf als schon beim früheren<br />

Zeitpunkt beobachtet (z. B. verringerte Expression von Histongenen), und Zeichen apoptotischer<br />

Prozesse wurden sichtbar (z. B. verstärkte Expression von BCL10, BLID, BAK1, BAG1).<br />

Das letzte Kapitel dieser Dissertation befaßt sich schließlich mit Tiling-Array Genexpressionsanalysen<br />

des Pathogens, die zunächst für S. aureus RN1HG in aeroben Schüttelkulturen für verschiedene<br />

Wachstumsphasen als Beginn und Referenzpunkt durchgeführt wurden. Anschließend<br />

wurde das bereits beschriebene in vitro Infektionsmodell der S9-Zellen eingesetzt, um Staphylokokken<br />

nach einer Phase der Internalisierung in ihren Wirtszellen zu den beiden Zeitpunkten<br />

2,5 h und 6,5 h nach Beginn der Infektion wieder zu extrahieren und ihre Genexpression zu<br />

charakterisieren. Das bakterielle Expressionspr<strong>of</strong>il zeigte die Aktivität des SaeRS Zwei-Komponenten-Systems<br />

in internalisierten Staphylokokken an. Zum Teil in Abhängigkeit von SaeRS wurde<br />

die stärkere Expression von Adhesinen (z. B. fnbAB, clfAB), Toxinen (hlgBC, lukDE, hla) und<br />

Genen, die der Umgehung des Immunsystems dienen (z. B. chp, eap), beobachtet. Außerdem<br />

wurden Expressionsveränderungen metabolischer Gene deutlich, z. B. verstärkte Expression von<br />

Genen der Aminosäurebiosynthese, des TCA-Zyklus und der Gluconeogenese. Dazu passend<br />

wurden Glycolysegene verringert exprimiert und des weiteren Gene der Purinbiosynthese und<br />

Gene, die tRNA-Synthetasen kodieren. Die Expressionsanalyse erfaßte ein bakterielles Genexpressionsprogramm,<br />

welches Literaturangaben einer spezifischen, von der Bakterienstamm-<br />

Wirtszellinien-Kombination abhängigen, transkriptionellen Adaptation des Erregers bestätigte.<br />

Neben klassischen physiologischen und mikrobiologischen Methoden tragen die modernen molekularbiologischen<br />

Technologien unter der Voraussetzung verfügbarer Genomsequenzen deutlich<br />

dazu bei, das Verständnis der Prozesse während des Zusammentreffens von Wirt und Pathogen<br />

zu verbessern. Auf der ersten Ebene der Reaktion verschaffen RNA-Expressionsuntersuchungen<br />

einen Überblick über die möglichen physiologischen und metabolischen Veränderungen. Aus<br />

diesem Grund sind die Ergebnisse eine wichtige Vervollständigung der Ergebnisse anderer Untersuchungsebenen<br />

und vergrößern die Grundlagenkenntnisse zum Geschehen während der<br />

Interaktion des Erregers mit seinem Wirt. Auf lange Sicht ist das Ziel dieser Grundlagenforschung,<br />

dazu beizutragen, Interventionsstrategien für Infektionskrankheiten zu etablieren.<br />

8


Maren Depke<br />

S U M M A R Y O F D I S S E R T A T I O N<br />

Humans and animals regularly encounter microorganisms like bacteria and need to defend<br />

themselves in order to avoid or limit damage. The <strong>host</strong> developed defense mechanisms to either<br />

avoid infection or to overcome it, whereas the <strong>pathogen</strong> gained mechanisms to evade these <strong>host</strong><br />

defense strategies. The <strong>host</strong>’s immune system is not only influenced <strong>by</strong> its own regulatory<br />

mechanisms and <strong>by</strong> the <strong>pathogen</strong>s’ factors, but also <strong>by</strong> additional elements like physical efforts<br />

or the psychological state <strong>of</strong> the organism. While short stressful episodes might even enhance<br />

the immune response, the immune response can be suppressed when the stress becomes<br />

chronic. But not only the immune system, but also metabolic processes might underlie<br />

modification <strong>by</strong> such stressors. Therefore, the improvement <strong>of</strong> knowledge on <strong>host</strong>-<strong>pathogen</strong><br />

<strong>interactions</strong> helps to elucidate mechanisms which benefit either <strong>host</strong> or <strong>pathogen</strong>. It will<br />

contribute to the establishment <strong>of</strong> new intervention strategies during infectious diseases. This<br />

thesis contains results from transcriptome studies on different aspects <strong>of</strong> <strong>host</strong>-<strong>pathogen</strong><br />

<strong>interactions</strong>.<br />

First, liver gene expression pr<strong>of</strong>iles from a murine chronic stress model served to elucidate<br />

aspects <strong>of</strong> the influence <strong>of</strong> stress on the metabolism and the immune response state <strong>of</strong> the<br />

animals. Psychological and physiological stressors can disturb neuroendocrine, immunological,<br />

behavioral, and metabolic functions and adaptive physiological processes aim to reconstitute a<br />

dynamic equilibrium. Recently, Kiank et al. described that BALB/c mice developed severe<br />

systemic immune suppression, neuroendocrinological disturbances and depression-like behavior<br />

due to 4.5 days <strong>of</strong> intermittent combined acoustic and restraint stress which served as a murine<br />

model <strong>of</strong> severe chronic psychological stress (Kiank et al. 2006; Brain Behav Immun. 20(4):359).<br />

Furthermore, mice subjected to chronic stress suffered from severe loss <strong>of</strong> body mass. Therefore,<br />

gene expression pr<strong>of</strong>iles <strong>of</strong> the liver, which plays the major role in metabolism, were analyzed to<br />

investigate primary causes for the observed loss <strong>of</strong> body weight. The liver fulfills an important<br />

function as a gatekeeper between the intestinal tract and the general circulation. Thus, the<br />

influence <strong>of</strong> psychological stress on immune regulatory processes in the liver was additionally<br />

investigated in the gene expression data set.<br />

Hepatic gene expression pr<strong>of</strong>iles <strong>of</strong> stressed mice exhibited distinct changes even after a<br />

single acute stress session. Although metabolic disturbances were not visible yet, a regulatory<br />

gene expression cascade started at that time point which led to the disturbances observed when<br />

stress had become chronic. Considering chronically stressed mice, genes related to metabolic<br />

diseases were most significantly influenced. Differential expression affected carbohydrate, amino<br />

acid, and lipid metabolism. Chronic stress in female BALB/c mice was shown to lead to a<br />

hypermetabolic syndrome including induction <strong>of</strong> gluconeogenesis, hypercholesteremia, and loss<br />

<strong>of</strong> essential amino acids. Additionally, the analysis <strong>of</strong> liver homogenates <strong>of</strong> acutely and<br />

chronically stressed mice revealed an altered mRNA expression <strong>of</strong> immune response genes.<br />

These included the induction <strong>of</strong> the acute phase response, but also <strong>of</strong> immune suppressive<br />

pathways. Furthermore, repression <strong>of</strong> hepatic antigen presentation was observed. In chronically<br />

stressed mice, increased leukocyte trafficking, increased oxidative stress together with counterregulatory<br />

gene expression changes, and an induction <strong>of</strong> apoptosis were detected.<br />

9


Maren Depke<br />

Summary <strong>of</strong> Dissertation<br />

The in vivo model <strong>of</strong> psychological stress in a complex mammalian <strong>host</strong> was performed<br />

without additional influences <strong>of</strong> a <strong>pathogen</strong>. This additional factor was addressed in the second<br />

study: Here, the influence <strong>of</strong> staphylococcal intra-venous infection on the <strong>host</strong> kidney gene<br />

expression was analyzed in another murine in vivo model using the wild type strain<br />

Staphylococcus aureus RN1HG and its isogenic sigB mutant.<br />

S. aureus, a Gram-positive bacterium, is a persistent commensal in the anterior nares <strong>of</strong><br />

approximately 20 % <strong>of</strong> the human population. The bacterium is found intermittently in 30 % to<br />

60 % <strong>of</strong> the population, and in non-carriers, which is the remaining fraction <strong>of</strong> the population,<br />

nasal swabs are never positive for S. aureus. Normally, such carriage does not cause any<br />

symptoms <strong>of</strong> illness. On the other hand, S. aureus can be responsible for a broad variety <strong>of</strong><br />

diseases: S. aureus can cause mild to severe local infections <strong>of</strong> skin or pharyngeal mucosa, but<br />

also <strong>of</strong> inner organs (e. g. endocarditis, osteomyelitis) and systemic disease like sepsis. S. aureus<br />

can be transmitted to the blood after body injury or <strong>by</strong> medical devices like catheters. An<br />

elementary model to mimic blood stream infection is the i. v. infection <strong>of</strong> laboratory animals, e. g.<br />

mice. Host reactions can be monitored <strong>by</strong> physiological, immunological or molecular readout<br />

systems. In this study, transcriptome analysis <strong>of</strong> murine kidney samples was performed.<br />

Although the virulence <strong>of</strong> sigB deficient strains is <strong>of</strong>ten reported to be similar to that <strong>of</strong> wild<br />

type strains the <strong>pathogen</strong>esis or pathomechanism <strong>of</strong> different infection settings might vary.<br />

Therefore, the rationale <strong>of</strong> this study was to investigate whether the deletion <strong>of</strong> sigB will lead to<br />

a different reaction <strong>of</strong> the infected <strong>host</strong>. Gene expression pr<strong>of</strong>iling indicated a highly<br />

reproducible <strong>host</strong> kidney response to infection with S. aureus. The comparison <strong>of</strong> infected with<br />

non-infected samples revealed a strong inflammatory reaction <strong>of</strong> kidney tissue. This included e. g.<br />

Toll-like receptor signaling, complement system, antigen presentation, interferon and IL-6<br />

signaling, but also counter-regulatory IL-10 signaling. However, the results <strong>of</strong> this study did not<br />

provide any hints for differences in the <strong>pathogen</strong>esis or pathomechanism <strong>of</strong> the S. aureus strains<br />

RN1HG and ΔsigB in the selected model <strong>of</strong> i. v. infection in mice, since the <strong>host</strong> response did not<br />

differ between infections with the two strains analyzed. If really existing, such differences might<br />

be transient and only apparent at earlier time points. Effects <strong>of</strong> SigB might also be compensated<br />

for in in vivo infection <strong>by</strong> the interlaced pattern <strong>of</strong> other regulators. There is also the possibility <strong>of</strong><br />

missing activity <strong>of</strong> SigB in vivo which could explain the similarity <strong>of</strong> <strong>host</strong> reaction to infection with<br />

S. aureus RN1HG and its sigB mutant in this study. SigB might possess only to a lesser extent<br />

characteristics attributed to virulence factors and might act in vivo more like a virulence<br />

modulator and fine tune bacterial reactions, or SigB might be important in special niches during<br />

infection. Assuming such function failure to detect differences in the <strong>host</strong>’s reaction to S. aureus<br />

RN1HG and its isogenic sigB mutant could be explained.<br />

Tissue expression pr<strong>of</strong>iling from in vivo models has the advantage <strong>of</strong> directly recording the<br />

relevant physiological state with all its complex <strong>interactions</strong> and influences and its vicinity to<br />

medical questions in the human. Nevertheless, it is very difficult to distinguish the different<br />

components because the tissue samples are always a mixture <strong>of</strong> different cell types which might<br />

even feature contrary reactions. Therefore, in vitro models were additionally analyzed in which<br />

only one defined <strong>host</strong> cell type was studied. Macrophages are an example for an immune cell<br />

type involved in the first steps <strong>of</strong> the encounter between the <strong>host</strong> and a <strong>pathogen</strong>. In the innate<br />

immune system, macrophages, together with dendritic cells, hold a central position. They are<br />

main effectors <strong>of</strong> the clearance <strong>of</strong> infections <strong>by</strong> their sentinel and phagocytic function.<br />

Macrophages present phagocytosed antigen derived peptides on MHC-II to lymphocytes. By this<br />

10


Maren Depke<br />

Summary <strong>of</strong> Dissertation<br />

function, macrophages take part in regulation <strong>of</strong> the adaptive immune response. The preparation<br />

<strong>of</strong> bone marrow stem cells and the in vitro differentiation <strong>of</strong> the stem cells into so-called bonemarrow<br />

derived macrophages (BMM) is a model to study reactions <strong>of</strong> macrophages. The<br />

advantage <strong>of</strong> this approach is that these macrophages have never been under any immunological<br />

influence which might result from the immune status <strong>of</strong> the animal even under standardized<br />

laboratory conditions. Until recently, BMM experiments were standardized only to a limited<br />

extent because serum-supplemented culture medium was used. To overcome sources <strong>of</strong><br />

experimental variation resulting from undefined and varying substances in serum, Eske et al.<br />

introduced serum-free culture conditions for BMM (Eske et al. 2009; J Immunol Methods. 342(1-<br />

2):13) which were now applied to study the reaction <strong>of</strong> BMM as third part <strong>of</strong> this thesis. BMM <strong>of</strong><br />

different mouse strains were treated with IFN-γ, a modulator <strong>of</strong> macrophage function, which is<br />

one <strong>of</strong> the first signals during initiation <strong>of</strong> the immune response in vivo. Host-<strong>pathogen</strong><br />

interaction experiments have revealed differences in reactions <strong>of</strong> BMM derived from BALB/c and<br />

C57BL/6 mice when confronted with Burkholderia pseudomallei especially after IFN-γ stimulation<br />

and at higher multiplicities <strong>of</strong> infection (Breitbach et al. 2006; Infect Immun. 74(11):6300). Also<br />

other infection studies uncovered differences between these two mouse strains in vivo and<br />

in vitro. Against the background <strong>of</strong> genetic influences on the BMM reactions, BALB/c and C57BL/6<br />

BMM were stimulated with IFN-γ to specify on a molecular level the reaction to the priming<br />

signal IFN-γ as basic principle. Furthermore, the study aimed to pr<strong>of</strong>ile potential differences <strong>of</strong><br />

reactions between the BMM <strong>of</strong> both mouse strains.<br />

Gene expression pr<strong>of</strong>iling revealed mainly induction <strong>of</strong> gene expression after treatment <strong>of</strong><br />

BMM with IFN-γ. Gene expression changes confirmed known IFN-γ effects like the induction <strong>of</strong><br />

immunoproteasome, antigen presentation and associated genes, interferon signaling related<br />

genes, GTPase/GTP binding protein genes, inducible nitric oxide synthase and others. IFN-γ<br />

dependent gene expression changes were highly similar in BALB/c and C57BL/6 BMM. Even for<br />

genes, which were differentially expressed only in BMM <strong>of</strong> one strain, a similar trend was<br />

observed in the other strain’s BMM. Gene expression differences between BMM <strong>of</strong> both strains<br />

were analyzed on the level <strong>of</strong> untreated controls as well as after IFN-γ treatment. Here,<br />

approximately 55 % to 60 % <strong>of</strong> the differentially expressed genes exhibited higher expression in<br />

BALB/c BMM, whereas the remaining genes featured higher expression in C57BL/6 BMM.<br />

Equivalent to the IFN-γ effects, a similar expression trend in the strain comparisons was visible at<br />

both treatment levels, even when regulation was significant only in one. Differentially expressed<br />

genes between BMM <strong>of</strong> both strains included immune-relevant genes as well as genes linked to<br />

cell death, but the coverage <strong>of</strong> functional groups was limited. The phenotypical differences<br />

between the reaction <strong>of</strong> BALB/c and C57BL/6 BMM were <strong>of</strong>ten determined in the presence <strong>of</strong><br />

IFN-γ and a second stimulus like LPS or infection. To elucidate molecular reasons for the observed<br />

differences in killing <strong>of</strong> <strong>pathogen</strong>s or cytokine production, the inclusion <strong>of</strong> samples subjected to a<br />

second stimulus in addition to IFN-γ is recommended.<br />

Not only immune cells or specially phagocytes get in touch with <strong>pathogen</strong>s, but also cells<br />

responsible for functional and structural integrity <strong>of</strong> <strong>host</strong> organs and tissue, like epithelial and<br />

endothelial cells. Such cells are actually part <strong>of</strong> the first line <strong>of</strong> recognition and reaction to a<br />

<strong>pathogen</strong>ic invasion into the <strong>host</strong>. While S. aureus colonizes humans in the anterior nares it can<br />

be a cause <strong>of</strong> pneumonia when transferred to the lung e. g. <strong>by</strong> aspiration or medical devices. The<br />

bronchial epithelial cell line S9 was used as an in vitro model system for the infection with<br />

staphylococci. A new experimental system permits the study <strong>of</strong> <strong>host</strong>-<strong>pathogen</strong> <strong>interactions</strong> in the<br />

11


Maren Depke<br />

Summary <strong>of</strong> Dissertation<br />

context <strong>of</strong> all bacterial factors, membrane-bound and secreted, and additionally prevents effects<br />

on bacterial physiology <strong>by</strong> prolonged handling, centrifugation and washing <strong>of</strong> bacteria (Schmidt<br />

et al. 2010; Proteomics. 10(15):2801). The fourth chapter in this thesis includes <strong>host</strong> gene<br />

expression signatures <strong>of</strong> S9 cells after in vitro infection with S. aureus RN1HG for the two time<br />

points 2.5 h and 6.5 h after start <strong>of</strong> infection. At the early time point, only the small number <strong>of</strong> 40<br />

genes was differentially expressed. Nevertheless, these few genes indicated a beginning proinflammatory<br />

response e. g. <strong>by</strong> the induction <strong>of</strong> cytokines (IL-6, IFN-β, LIF) or prostaglandinendoperoxide<br />

synthase 2 (PTGS2), but also counter-regulatory processes were discernible e. g. <strong>by</strong><br />

the induction <strong>of</strong> CD274, ligand for the immunoinhibitory receptor PD-1. Furthermore, negative<br />

effects on cell cycle progression became visible. The <strong>host</strong> cell response was dramatically<br />

aggravated at the later 6.5 h time point. Differential expression was detected for 1196 genes.<br />

These included induced cytokines, pattern recognition receptor signaling, antigen presentation,<br />

and genes involved in immune defense (e. g. GBPs, MX, APOL). Negative effects on growth and<br />

proliferation were even more enhanced in comparison to the early time point, e. g. repression <strong>of</strong><br />

histone genes, and signs for apoptotic processes were revealed, e. g. induction <strong>of</strong> BCL10, BLID,<br />

BAK1, and BAG1.<br />

Finally, the last chapter addresses the <strong>pathogen</strong>’s expression pr<strong>of</strong>ile, which was first recorded<br />

from agitated, aerobic S. aureus RN1HG cultures in different growth phases as a starting and<br />

reference point. Afterwards, the already described S9 cell in vitro infection model was used to<br />

extract staphylococci after an internalization phase inside the <strong>host</strong> cell. Internalized bacteria<br />

were analyzed at the two time points 2.5 h and 6.5 h after start <strong>of</strong> infection in comparison to<br />

different control samples <strong>by</strong> tiling array gene expression analysis.<br />

Pathogen expression pr<strong>of</strong>iling revealed the activity <strong>of</strong> the SaeRS two-component system in<br />

internalized staphylococci. Partly dependent on this regulatory system, the induction <strong>of</strong> adhesins<br />

(e. g. fnbAB, clfAB), toxins (hlgBC, lukDE, hla), and immune evasion genes (e. g. chp, eap) was<br />

observed. Furthermore, expression changes <strong>of</strong> metabolic genes were recorded. This included the<br />

induction <strong>of</strong> amino acid biosynthesis pathways, TCA cycle genes, and gluconeogenesis. In<br />

accordance with these findings, glycolysis genes were repressed as well as purine biosynthesis<br />

and tRNA synthetase genes. Expression analysis recorded a distinct bacterial expression program,<br />

which supported literature results <strong>of</strong> a specific, bacterial strain and <strong>host</strong> cell line dependent<br />

transcriptional adaptation <strong>of</strong> the <strong>pathogen</strong>.<br />

Different approaches can be applied to define even more and specific <strong>interactions</strong> between<br />

the <strong>host</strong> and the <strong>pathogen</strong>. Besides classical physiological and microbiological methods, the<br />

modern molecular technologies can considerably help to improve the understanding <strong>of</strong> the<br />

processes acting during encounter <strong>of</strong> <strong>host</strong> and <strong>pathogen</strong> provided that genome sequence<br />

information is available. On the first level <strong>of</strong> reaction, RNA expression pr<strong>of</strong>iling will generate an<br />

overview on the potential changes <strong>of</strong> physiology and metabolism. Therefore, the results are an<br />

important complementation <strong>of</strong> results from other analysis levels and increase the fundamental<br />

knowledge on processes during the encounter <strong>of</strong> <strong>pathogen</strong> and its <strong>host</strong>. In the long-term, this<br />

basic research will probably contribute to the establishment <strong>of</strong> intervention strategies during<br />

infectious disease.<br />

12


Maren Depke<br />

I N T R O D U C T I O N<br />

INFECTIOUS DISEASES AND IMMUNE SYSTEM<br />

Humans and animals regularly encounter microorganisms like bacteria and need to defend<br />

themselves in order to avoid or limit damage <strong>of</strong> their own organismal integrity.<br />

Before any possibly dangerous interaction could occur, they first aim to prevent the nearer<br />

contact <strong>by</strong> physical mechanisms. Dry, closely connected skin surfaces impede the settlement <strong>of</strong><br />

microorganisms, and inner surface epithelium <strong>of</strong>ten uses mucus as trapping and rinsing agent or<br />

creates unfavorable conditions e. g. <strong>by</strong> low pH in the stomach. Such physical barriers cannot<br />

completely prevent colonization. A commensalism <strong>of</strong> certain microorganisms has been<br />

established. But this is not a disadvantage: The occupation <strong>of</strong> interaction surfaces <strong>by</strong> a<strong>pathogen</strong>ic<br />

commensals limits the available habitat for <strong>pathogen</strong>s.<br />

Additionally, the <strong>host</strong> organism has developed active defense mechanisms. On the one hand,<br />

general defense factors are produced in advance and deposited at potential interaction sites.<br />

Such factors, like lytic enzymes or defensin peptides, will exert their properties without further<br />

assistance <strong>of</strong> the <strong>host</strong>’s physiological systems. These factors can additionally be further induced<br />

after initiation <strong>of</strong> the immune response. On the other hand in case <strong>of</strong> a successful infection <strong>by</strong> a<br />

<strong>pathogen</strong>, the <strong>host</strong> is able to start reaction cascades, which will be activated and enhanced after<br />

contact with the <strong>pathogen</strong>. The <strong>host</strong> possesses receptors which bind conserved structures <strong>of</strong><br />

<strong>pathogen</strong>s. This enables the <strong>host</strong> to recognize the infection, initialize alarm cascades, and<br />

activate further antimicrobial factors, which are prepared as inactive precursors. The <strong>host</strong> also<br />

activates the blood coagulation cascade and <strong>by</strong> this means aims to confine infection to specific<br />

sites and to reduce spreading. Also cellular defense processes are initiated in sequence <strong>of</strong><br />

<strong>pathogen</strong> infiltration into the body, where a first step is <strong>pathogen</strong>-unspecific phagocytosis <strong>by</strong><br />

macrophages or neutrophils. In the later phases <strong>of</strong> infection, the <strong>host</strong> can adapt even more<br />

specialized defense systems to react to the specific infecting agent on a humoral and cellular<br />

level. This approach needs more time, but is most effective and finally enables the <strong>host</strong> to<br />

overcome the infectious disease.<br />

Aspects <strong>of</strong> Innate Immunity<br />

Many <strong>pathogen</strong>s possess conserved structures. These can be related to their outer structure<br />

like peptidoglycan <strong>of</strong> the bacterial cell wall, but also to the very basic genetic information like the<br />

double-stranded RNA constituting the genome <strong>of</strong> some viruses. Some <strong>of</strong> them are essential for<br />

the <strong>pathogen</strong> and therefore highly conserved during evolution. In their co-evolution, both sides,<br />

13


Maren Depke<br />

Introduction<br />

<strong>host</strong> as well as <strong>pathogen</strong>, developed mechanisms to fight or to gain some advantage against the<br />

other. The <strong>host</strong> developed defense mechanisms to either avoid infection or to overcome it,<br />

whereas the <strong>pathogen</strong> gained mechanisms to evade these <strong>host</strong> defense strategies. One part <strong>of</strong><br />

the <strong>host</strong>’s immune defense is the innate immune system, for which the detailed information is<br />

encoded in the genome and which has been passed on and further modified and optimized<br />

during evolution. The <strong>host</strong> as a complex organism commands two main levels <strong>of</strong> such immune<br />

defense, non-cellular and cellular. On epithelial barriers, the <strong>host</strong> deposits antimicrobial peptides,<br />

which are <strong>of</strong>ten membrane-active and pore-forming and aim to lyse the <strong>pathogen</strong> during its<br />

attempt to enter the body (Schröder 1999). Also enzymes like lyzozyme are secreted and serve a<br />

similar function.<br />

A very important and complex defense system <strong>of</strong> cascade-activated components is the<br />

complement system (Dunkelberger/Song 2010). The complement system consists <strong>of</strong> several<br />

serum proteins, which interact and build complexes, contain several sequential and<br />

interdependent activation steps and enzymatic activities, and finally achieves three main goals:<br />

First, in a process called opsonization the labeling <strong>of</strong> <strong>pathogen</strong>s with a molecular tag to render it<br />

recognizable for phagocytosis, and second the lysis <strong>of</strong> the <strong>pathogen</strong> <strong>by</strong> pore-forming<br />

components. The third function during activation <strong>of</strong> the complement is to pass the information <strong>of</strong><br />

occurring infection to the other immune defense systems.<br />

Three activation pathways <strong>of</strong> the complement system exist (Fig. I.1). The first, called classical<br />

pathway, is linked to the non-innate, i. e. adaptive immune system, because it relies in the first<br />

step on antibodies, which recognize specific <strong>pathogen</strong>ic structures. Pathogen surface bound<br />

antibodies bind and activate the C1 factor <strong>of</strong> the complement system. The following steps <strong>of</strong><br />

proteolytic cleavage produce from factors C2 and C4 the smaller “a” and the bigger “b”<br />

fragments. C2a and C4b together form the first very important complex during complement<br />

activation, the C3-convertase, which cleaves C3 into C3a and C3b. The C3-convertase (via C4b)<br />

and C3b will be covalently bound to the <strong>pathogen</strong>’s cell surface after reaction <strong>of</strong> an active<br />

thioester bond and accomplish opsonization.<br />

The second complement activation pathway leads to the same effects, but starts with a<br />

different recognition complex including lectins (Fujita et al. 2004). This lectin-activation pathway<br />

does not require antibodies bound to the surface <strong>of</strong> the <strong>pathogen</strong>. Here, mannose-binding lectin<br />

(MBL) or ficolin recognize conserved carbohydrate structures <strong>of</strong> the <strong>pathogen</strong>. These molecules<br />

belong to the so-called pattern recognition receptors (PRRs), conserved and non-variable<br />

receptors for conserved <strong>pathogen</strong>ic structures, which in turn are called <strong>pathogen</strong>-associated<br />

molecular patterns (PAMPs). MBL or ficolin are complexed, among others, with a serine protease<br />

(MASP), which achieves the next activation steps <strong>of</strong> C2 and C4 and formation <strong>of</strong> the C3-<br />

convertase as described before in the classical pathway.<br />

The third way <strong>of</strong> complement activation is named alternative pathway. For activation, it uses<br />

the spontaneous hydrolysis <strong>of</strong> C3 in the plasma to C3(H 2 O), which exposes the molecular site for<br />

covalent binding to <strong>pathogen</strong>ic surfaces. In the presence <strong>of</strong> such surface, C3(H 2 O) binds to it and<br />

subsequently complement factor B to C3(H 2 O). Finally, factor D cleaves factor B into the<br />

fragments Ba and Bb. The resulting complex <strong>of</strong> C3(H 2 O)Bb constitutes the initial C3-convertase <strong>of</strong><br />

the alternative pathway and cleaves further C3 molecules. The resulting C3b fragments again<br />

opsonize the <strong>pathogen</strong>, but also bind factor B and D, and form further C3-convertase complexes<br />

C3bBb. At this step, a strong amplification takes place, also when the C3b fragments are derived<br />

from activation <strong>of</strong> the classical and lectin pathway.<br />

14


Maren Depke<br />

Introduction<br />

Besides opsonization, the complement system aims to lyse the <strong>pathogen</strong>. This so-called<br />

effector function is initiated after formation <strong>of</strong> C3b <strong>by</strong> any <strong>of</strong> the activation pathways. When C3b<br />

is complexed with the C3-convertases <strong>of</strong> classical/lectin or alternative pathway it forms the C5-<br />

convertase complex (C2bC4bC3b or C3bBbC3b). The enzymatic function is the cleavage <strong>of</strong><br />

complement factor C5 into C5b and C5a. Now, C5b initiates the formation <strong>of</strong> the membrane<br />

attack complex (MAC). Sequential binding <strong>of</strong> factors C6, C7 (membrane binding) and C8<br />

(membrane intrusion) leads to the forming <strong>of</strong> a pore <strong>by</strong> several C9 molecules. Such pores finally<br />

can lead to the lysis <strong>of</strong> the <strong>pathogen</strong>. Because the complement system is a very powerful defense<br />

system, which could lead to immense damage to the <strong>host</strong> itself, it is tightly controlled <strong>by</strong> many<br />

regulatory proteins which especially prevent the activation on the <strong>host</strong>’s own cellular structures.<br />

Fig. I.1: Complement activation pathways.<br />

Abbreviations: MBL – mannose-binding lectin; MASP – MBL-associated serine protease; sMAP – small MBL-associated protein;<br />

C3(H 2O) – hydrolysed C3<br />

From: Foster 2005.<br />

During the activation <strong>of</strong> the complement cascade, several small peptide fragments, called<br />

anaphylatoxins, are produced. C3a, C5a, and to a lesser extent C4a have inflammatory properties.<br />

In a localized infection, they lead to increased permeability <strong>of</strong> blood vessels, induce endothelial<br />

adhesion molecules, influence monocytes and neutrophils to increase attachment, and activate<br />

mast cells which further enforce the effect with their own mediators.<br />

Hence, the small peptides link the non-cellular to the cellular part <strong>of</strong> the innate immune<br />

system. Phagocytes have a central function in the innate immune defense. These cells clear the<br />

<strong>pathogen</strong>s from the site <strong>of</strong> infection and kill them intracellularly. Monocytes from the blood<br />

stream migrate into the tissue and develop into long-living macrophages, which reside and are<br />

activated in case <strong>of</strong> an encounter with <strong>pathogen</strong>s. Several organs, especially those at a higher risk<br />

to be exposed to <strong>pathogen</strong>s, embody special types <strong>of</strong> macrophages, e. g. Kupffer cells <strong>of</strong> the liver.<br />

15


Maren Depke<br />

Introduction<br />

Neutrophil granulocytes are another group <strong>of</strong> phagocytes. They are found in the blood stream<br />

and leave it only at sites <strong>of</strong> infection. During an infection, the numbers <strong>of</strong> neutrophils can<br />

increase strongly, but these cells are only short-living. Phagocytes recognize <strong>pathogen</strong>ic<br />

structures <strong>by</strong> membrane PRRs. Binding to the <strong>pathogen</strong> initiates phagocytosis. First, the<br />

<strong>pathogen</strong> is enclosed in the phagosomal compartment, which afterwards fuses to lysosomes.<br />

Lysosomes bring enzymes and antimicrobial mediators to the new phagolysosomal compartment<br />

and finally accomplish killing <strong>of</strong> the <strong>pathogen</strong> in a process called respiratory burst, during which<br />

enzymes produce toxic reaction products like H 2 O 2 , O 2 – , and NO under consumption <strong>of</strong> O 2 , the<br />

name-giving effect. During activation, phagocytes produce cytokines and chemokines, which lead<br />

to a pro-inflammatory reaction and to chemotaxis <strong>of</strong> further immune cells like monocytes and<br />

neutrophils (Janeway et al. 2002).<br />

Aspects <strong>of</strong> Adaptive Immunity<br />

A third type <strong>of</strong> phagocytic cells, dendritic cells, links the innate to the adaptive immune<br />

system. These cells ingest different antigens in the peripheral tissue <strong>by</strong> phagocytosis and<br />

macropinocytosis and and transport them to the lymph nodes, where non-activated cells <strong>of</strong> the<br />

apaptive immune response wait for an activation trigger. Dendritic cells are the mediator cells<br />

which relay the information <strong>of</strong> peripherally present antigens and therefore potential infection to<br />

cells which might bear the specific receptor to these antigens. The information is transferred in a<br />

process called antigen presentation. It involves major histocompatibility complex (MHC)<br />

molecules. The complex is formed either <strong>by</strong> an α-chain with transmembrane domain and a noncovalently<br />

linked beta-2-microglobulin (B2M) molecule as β-chain in class I type <strong>of</strong> MHC or <strong>by</strong> an<br />

α- and a β-chain which are both inserted in the membrane in class II type <strong>of</strong> MHC (Fig. I.2 A).<br />

MHC-I α-chains and MHC-II α- and β-chains form a binding groove for antigen derived peptides<br />

and are encoded in the genome in a highly polymorph manner. Only B2M is encoded <strong>by</strong> a single<br />

gene. The two types <strong>of</strong> this complex, class I and class II, present antigenic peptides from different<br />

origin to subsets <strong>of</strong> T cells, which in turn possess a receptor for MHC-peptide-complexes, the<br />

T cell receptor (TCR).<br />

A<br />

Fig. I.2:<br />

Antigen presentation to T cells via MHC molecules.<br />

A. T cell receptor (TCR) mediated recognition <strong>of</strong> a peptide antigen bound<br />

to an MHC-derived molecule.<br />

B. Activation <strong>of</strong> naïve T helper cells requires the presence <strong>of</strong> co-stimulatory<br />

molecules, e. g. B7.<br />

Abbreviations:<br />

MHC – major histocompatibility complex; APC – antigen presenting cell<br />

From: Andersen et al. 2006.<br />

B<br />

16


Maren Depke<br />

Introduction<br />

Extracellular antigens are presented <strong>by</strong> MHC-II molecules. After phagocytosis, these antigens<br />

are degraded in the phagolysosome. Transport vesicles from the Golgi, which carry unloaded<br />

MHC-II complexes, fuse to the phagolysosome, peptides are loaded to the MHC-II, and the<br />

complete complex is transferred to the cell surface.<br />

MHC-I serves the presentation <strong>of</strong> intracellular antigens. Here, the cleavage <strong>of</strong> proteins is<br />

performed in the cytosol <strong>by</strong> the proteasome, a high molecular weight protease complex. Peptide<br />

transporters transfer the peptides to the ER, where the loading <strong>of</strong> MHC-I molecules takes place.<br />

Antigen presentation via MHC-II is restricted to pr<strong>of</strong>essional antigen presenting cells (APC) <strong>of</strong><br />

the immune system, whereas presentation <strong>of</strong> intracellular antigens via MHC-I is additionally<br />

present in many non-immune cells. Interestingly, natural killer (NK) cells, which belong to the<br />

innate immune response and do not possess antigen specific receptors, sense the missing <strong>of</strong><br />

MHC-I presentation, which occurs in some virus-infected or cancer cells, and combat these cells<br />

<strong>by</strong> inducing apoptosis (Jensen 1999).<br />

The TCR <strong>of</strong> T cells is only able to bind MHC-peptide complexes tightly when the TCR features<br />

the corresponding specificity to the peptide antigen. A very efficient system uses combination <strong>of</strong><br />

different TCR chains and somatic recombination, which is the removal <strong>of</strong> certain genomic<br />

sections and recombining <strong>of</strong> the remaining parts, to generate an immense repertoire <strong>of</strong> different<br />

T cell clones with different antigen specificity. These cells are additionally selected for nonreactivity<br />

to the <strong>host</strong>’s own antigens and build a reservoir <strong>of</strong> defense cells for <strong>pathogen</strong>ic<br />

antigens which might interfere with the <strong>host</strong> during life-time. This is the distinctive feature <strong>of</strong> the<br />

adaptive immune system: It carries the potential for the defense against almost every <strong>pathogen</strong>,<br />

but it realizes in an adaptive manner with high specificity only the defense needed in the really<br />

occurring case <strong>of</strong> infection, which makes it a time-consuming, but highly effective defense<br />

mechanism against <strong>pathogen</strong>s.<br />

T-cells are subdivided into populations differing <strong>by</strong> the expression <strong>of</strong> surface antigens and <strong>by</strong><br />

the secreted cytokines, which in summary links them to distinct functions. The main two groups<br />

are characterized <strong>by</strong> expression <strong>of</strong> CD4 or CD8. CD8 + T cells are cytotoxic effector cells. Their TCR<br />

binds MHC-I molecules and therefore is able to recognize <strong>pathogen</strong>ic intracellular antigens.<br />

Specific binding in case <strong>of</strong> intracellular infection activates processes to destroy the cell and kill the<br />

<strong>pathogen</strong>. CD4 + T cells are also called helper T cells. Their TCR binds to MHC-II molecules and thus<br />

recognizes antigens derived after phagocytosis. In consequence, CD4 + T cells initiate mechanisms<br />

fighting extracellular infection or disposing extracellular antigens. This is on the one hand a link<br />

back to the innate immune cells, e. g. macrophages, which can be activated, but on the other a<br />

link to a further aspect <strong>of</strong> the adaptive immune response. Besides the cellular adaptive immune<br />

response the <strong>host</strong> commands an adaptive humoral immunity mediated <strong>by</strong> antibodies, which<br />

originate from B cells (plasma B cells). In a process similar to the generation <strong>of</strong> specific TCR in<br />

T cells, the B cell produces its own B cell receptor (BCR), which already includes the specificity <strong>of</strong><br />

the later antibodies. B cells are able to bind antigens via their BCR, phagocytose and degrade<br />

them and finally present them on their surface as MHC-II-peptide complexes. An activated CD4 +<br />

T cell with a specificity to this peptide can activate the B cell, which in turn starts further<br />

differentiation steps and the production <strong>of</strong> antibodies. These antibodies autonomously bind their<br />

specific antigen, and constitute an opsonization factor for the clearance <strong>of</strong> antibody-antigen<br />

complexes <strong>by</strong> phagocytosis. Furthermore, they are the activation complex for the complement<br />

system as described before (Janeway et al. 2002).<br />

17


Maren Depke<br />

Introduction<br />

As T cells have a central position in the immune defense as cytotoxic effectors or potent<br />

activators <strong>of</strong> other immune cells, MHC-molecule binding alone does not lead to their activation,<br />

but co-stimulatory signals are necessary. These signals originate from co-stimulatory surface<br />

receptors like the B7-family <strong>of</strong> proteins expressed <strong>by</strong> antigen-presenting cells (APC), e. g. the<br />

dendritic cells. Only when both signals initiate signal transduction at the same time, the T cell<br />

activation takes place (Fig. I.2 B).<br />

Modulation <strong>of</strong> Immune Reactions<br />

The immune system is not only influenced <strong>by</strong> its own regulatory mechanisms and <strong>by</strong> the<br />

<strong>pathogen</strong>s’ factors, but also <strong>by</strong> additional elements like physical efforts (Gleeson et al. 1995) or<br />

the psychological state <strong>of</strong> the organism (Glaser et al. 1999). While short stressful episodes might<br />

even enhance the immune response, the immune response can be suppressed when the stress is<br />

lasting too long. Immune suppression is mainly mediated <strong>by</strong> glucocorticoids (Dallman 2007).<br />

Thus, different stressors might affect the demands <strong>of</strong> time for fighting an infection or even the<br />

success <strong>of</strong> immune defense mechanisms (Fig. I.3, Peterson PK et al. 1991, West et al. 2006). But<br />

not only the immune system, but also metabolic processes might underlie modification <strong>by</strong> such<br />

stressors. Increasing demands and overwhelming environmental stimuli in the modern society<br />

continuously heighten the stress level <strong>of</strong> humans and escalate the <strong>pathogen</strong>esis <strong>of</strong> stressassociated<br />

illness such as the metabolic syndrome or depression and increase the risk <strong>of</strong><br />

infections (Bartolomucci 2007, Leonard 2006, Lundberg 2005).<br />

STRESSOR<br />

type<br />

intensity<br />

timing<br />

duration<br />

PATHOGEN<br />

species, strain (virulence)<br />

inoculum size<br />

HOST<br />

species (genetics)<br />

age<br />

sex<br />

concomitant disease<br />

nutritional status<br />

previous experience<br />

with <strong>pathogen</strong> (immunity)<br />

with stressor (tolerance)<br />

INFECTIOUS DISEASE<br />

Fig. I.3: Various factors can modify the impact <strong>of</strong> a stressor on the<br />

<strong>pathogen</strong>esis <strong>of</strong> an infectious disease.<br />

From: Peterson et al. 1991.<br />

symptomatic<br />

infection<br />

DEATH<br />

asymptomatic<br />

infection<br />

HEALTH<br />

A physiological stress response is short lasting and physiologically important for survival to<br />

cope with a changing environment or to deal with potentially life-threatening situations.<br />

Adaptive processes are very rapidly mounted to reconstitute a balanced allostatic system in the<br />

stressed body which primarily include the neuroendocrine and immune system. Stress-induced<br />

neuroendocrine alterations include activation <strong>of</strong> the sympathetic nervous system with increased<br />

secretion <strong>of</strong> catecholamines, and stimulation <strong>of</strong> the hypothalamus-pituitary-adrenal (HPA) axis<br />

18


Maren Depke<br />

Introduction<br />

with heightened release <strong>of</strong> glucocorticoids. Prolonged and increased release <strong>of</strong> catecholamines is<br />

associated with cardiovascular diseases such as hypertension, myocardial infarction, or stroke.<br />

Excessive secretion <strong>of</strong> glucocorticoids was linked to diabetes, dyslipidemia, cardiovascular<br />

alterations, immunosuppression, and mood disorders (Lundberg 2005, McEwen 2004,<br />

Viswanathan/Dhabhar 2005, Wrona 2006).<br />

Recently, Kiank et al. described that BALB/c mice developed severe systemic immune<br />

suppression, neuroendocrinological disturbances and depression-like behavior due to 4.5 days <strong>of</strong><br />

intermittent combined acoustic and restraint stress which serves as a murine model <strong>of</strong> severe,<br />

chronic psychological stress (Kiank et al. 2006, 2007a). Immunosuppression was substantiated <strong>by</strong><br />

a heightened anti-inflammatory cytokine bias, an apoptotic loss <strong>of</strong> lymphocytes leading to<br />

lymphocytopenia, T cell anergy, and impaired phagocytic and oxidative burst responses. The<br />

immunodeficient state increased the animals’ susceptibility to experimental infection with E. coli<br />

(Kiank et al. 2006, 2007b). Repeatedly stressed mice actually developed spontaneous bacterial<br />

infiltrations into the lung measurable even 7 days after the last chronic stress cycle, which was<br />

associated with a reduced inducibility <strong>of</strong> IFN-γ, a cytokine that was shown to be important to<br />

prevent spreading <strong>of</strong> translocated commensals from the gut (Kiank et al. 2008). On the one hand,<br />

stress-induced immunosuppression was accompanied with a reduced clearance <strong>of</strong> experimental<br />

infections in the long term, but on the other hand, with attenuation <strong>of</strong> a hyperinflammatory<br />

septic shock (Kiank et al. 2006, 2007a). Furthermore, behavioral alterations with increased<br />

depression-like behavior and neuroendocrine alterations such as prolonged activation <strong>of</strong> the HPA<br />

axis and increased turnover <strong>of</strong> catecholamines were documented.<br />

Finally, a prominent stress-induced loss <strong>of</strong> body mass without significant changes <strong>of</strong> food and<br />

water intake during the observation period became detectable (Kiank et al. 2006). It is known<br />

that stress exposure is linked to changes <strong>of</strong> body weight. There is evidence that hypothalamic<br />

control <strong>of</strong> food intake is influenced <strong>by</strong> stress, which in consequence alters metabolism. In such a<br />

situation, some people lose and others gain weight in response. However, the molecular<br />

mechanisms <strong>of</strong> the stress to body weight connection remain to be elucidated.<br />

It is important for the understanding <strong>of</strong> stress-induced immunoregulatory mechanisms that<br />

stress hormones like catecholamines and glucocorticoids can modulate several liver functions<br />

including carbohydrate, protein and lipid metabolism or affect the immune response. Especially<br />

corticosteroids can suppress inflammatory processes <strong>by</strong> preventing the release <strong>of</strong><br />

proinflammatory mediators, <strong>by</strong> diminishing immune cell trafficking, phagocytosis and radical<br />

production or <strong>by</strong> down-regulating antigen presentation, <strong>by</strong> inhibiting lymphocyte proliferation,<br />

and <strong>by</strong> inducing apoptosis <strong>of</strong> immune cells. Thus, prolonged increase <strong>of</strong> glucocorticoid levels<br />

during chronic stress can activate hepatic catabolic pathways but also effectively suppress the<br />

local immune response (Bartolomucci 2007, Chida et al. 2006, Elenkov 2004, Swain 2000).<br />

19


Maren Depke<br />

Introduction<br />

STAPHYLOCOCCUS AUREUS<br />

General Features <strong>of</strong> S. aureus<br />

Staphylococci are Gram-positive bacteria <strong>of</strong> approximately 1 µm diameter, which <strong>of</strong>ten<br />

aggregate in grape-like structures. The golden color <strong>of</strong> Staphylococcus aureus colonies – derived<br />

from staphyloxanthin and other C 30 triterpenoid carotenoids (Marshall/Wilmoth 1981) which<br />

function as antioxidant protection molecules (Liu GY et al. 2005) – led to its naming. Several<br />

S. aureus properties like the ability to perform hemolysis and expression <strong>of</strong> catalase and<br />

coagulase allow differentiation <strong>of</strong> the species from other bacteria. S. aureus can produce energy<br />

<strong>by</strong> aerobic respiration, but is facultatively able to survive in anaerobic environments. In situations<br />

without oxygen it can apply lactate fermentation. S. aureus seems to be naturally auxotrophic<br />

and requires e. g. amino acids as growth supplement. The staphylococcal cell wall peptidoglycan<br />

features a special crossbridge structure formed <strong>by</strong> multiple glycine residues. These are the target<br />

structures for lysostaphin, a staphylocidal enzyme from Staphylococcus simulans first described<br />

<strong>by</strong> Schindler and Schuhardt in 1964, which was taken over <strong>by</strong> scientist for several laboratory<br />

research purposes and which might even be applied as mupirocin substitute for eradication <strong>of</strong><br />

staphylococcal nasal colonization (Recsei et al. 1987, Kokai-Kun et al. 2003). Staphylococci are<br />

resistant to lysozyme, a muramidase, which is present as antimicrobial enzyme in body fluids and<br />

at sites <strong>of</strong> infection as part <strong>of</strong> the innate immune defense. The resistance is mediated <strong>by</strong> O-<br />

acetylation in the muramic acid <strong>of</strong> peptidoglycan (position C6-OH) which is conducted <strong>by</strong> the<br />

peptidoglycan-specific O-acetyltransferase OatA. OatA mutants are susceptible to lysozyme (Bera<br />

et al. 2005).<br />

The S. aureus genome has 32 % GC content. Its chromosome is about 2.7E+06 to 2.9E+06<br />

nucleotides in length, contains about 2600 to 3000 genes, and has been sequenced for several<br />

strains until now (Table I.1). Additionally, sequence information for several plasmids is available<br />

(http://www.ncbi.nlm.nih.gov/sites/entrez?db=Genome).<br />

For sub-division purposes, the S. aureus genome has been classified into two parts. The core<br />

genome, which occupies approximately 75 % <strong>of</strong> the complete genome, includes genes with basic<br />

functions e. g. in house-keeping and central metabolism. These genes are present in the genomes<br />

<strong>of</strong> most strains and exhibit a high degree <strong>of</strong> conservation in the individual genes’ sequences in<br />

different strains. The core part also features a high similarity to the phylogenetically related<br />

Bacillus species genomes (Hiramatsu et al. 2004). The remaining 25 % belong to the accessory<br />

part <strong>of</strong> the genome. This part consists <strong>of</strong> further genes which have been acquired <strong>by</strong> some<br />

S. aureus strains e. g. from mobile genetic elements. Here, variation in the presence <strong>of</strong> genes is<br />

observed in different strains. Many virulence genes are classified as fraction <strong>of</strong> the accessory<br />

genome (Lindsay/Holden 2004).<br />

S. aureus has many options <strong>of</strong> varying its accessory genome part <strong>by</strong> different mobile genetic<br />

elements (Plata et al. 2009). Staphyloccoccal <strong>pathogen</strong>icity islands (SaPI) are located at constant<br />

positions <strong>of</strong> the genome, contain superantigen genes and sequences with similarity to prophages<br />

(Lindsay et al. 1998, Novick RP 2003). Similarly, genes <strong>of</strong> the νSa-family <strong>of</strong> <strong>pathogen</strong>icity genomic<br />

islands, <strong>of</strong> which several types (1-4, α, β, γ) exist, code for virulence factors and toxins (Gill et al.<br />

20


Maren Depke<br />

Introduction<br />

2005). SaPIs and some νSa-members can be excised from the genome and are involved in<br />

horizontal gene transfer. S. aureus strains <strong>of</strong>ten harbor prophages, which also incorporate<br />

virulence factors and are an aid for their transfer between strains and in general influence<br />

S. aureus evolution (Lindsay/Holden 2004). The Panton-Valentine leukocidin is a phage-encoded<br />

staphylococcal toxin. S. aureus NCTC8325 carries three lysogenic phages, φ11, φ12, and φ13, <strong>of</strong><br />

which φ13 introduces the virulence factor staphylokinase (sak) into the strain (Iandolo et al.<br />

2002). Kwan et al. reported in 2005 the genome sequences and predicted proteins <strong>of</strong> 27<br />

bacteriophages <strong>of</strong> S. aureus (Kwan et al. 2005). Additionally, gene transfer can be accomplished<br />

<strong>by</strong> shorter insertion sequences or <strong>by</strong> longer transposons which bring along their own transposase<br />

genes and recognition sites (Hiramatsu et al. 2004, Plata et al. 2009). Plasmids are a further place<br />

<strong>of</strong> encoding useful, although not essential information, e. g. for antibiotic resistances. S. aureus<br />

strains include different plasmids <strong>of</strong> varying size and copy number. For example, the methicillinresistant<br />

S. aureus strain COL owns the 4440 nt plasmid pT181 (accession number NC_006629;<br />

http://www.ncbi.nlm.nih.gov/sites/entrez?db=Genome), whose sequence became available in<br />

2005. Some S. aureus strains have also acquired the vanA operon for vancomycin resistance from<br />

an Enterococcus spp. transposon via a conjugative plasmid (Péricon/Courvalin 2009).<br />

Table I.1: Complete, RefSeq genome sequences for bacterial chromosomes <strong>of</strong> Staphylococcus aureus strains from NCBI Entrez<br />

Genome database (http://www.ncbi.nlm.nih.gov/sites/entrez?db=Genome) <strong>of</strong> August 25 th 2010.<br />

accession number strain length / nt date created genes<br />

NC_002745 Staphylococcus aureus subsp. aureus N315 2814816 2001/04/21 2664<br />

NC_002951 Staphylococcus aureus subsp. aureus COL 2809422 2001/09/14 2723<br />

NC_002758 Staphylococcus aureus subsp. aureus Mu50 2878529 2001/10/04 2774<br />

NC_002952 Staphylococcus aureus subsp. aureus MRSA252 2902619 2001/11/06 2839<br />

NC_002953 Staphylococcus aureus subsp. aureus MSSA476 2799802 2001/11/06 2715<br />

NC_003923 Staphylococcus aureus subsp. aureus MW2 2820462 2002/05/31 2704<br />

NC_007622 Staphylococcus aureus RF122 2742531 2005/11/24 2663<br />

NC_007793 Staphylococcus aureus subsp. aureus USA300_FPR3757 2872769 2006/02/11 2648<br />

NC_007795 Staphylococcus aureus subsp. aureus NCTC8325 2821361 2006/02/18 2969<br />

NC_009487 Staphylococcus aureus subsp. aureus JH9 2906700 2007/05/23 2816<br />

NC_009632 Staphylococcus aureus subsp. aureus JH1 2906507 2007/07/03 2870<br />

NC_009641 Staphylococcus aureus subsp. aureus Newman 2878897 2007/07/06 2687<br />

NC_009782 Staphylococcus aureus subsp. aureus Mu3 2880168 2007/09/06 2768<br />

NC_010079 Staphylococcus aureus subsp. aureus USA300_TCH1516 2872915 2007/12/03 2799<br />

NC_013450 Staphylococcus aureus subsp. aureus ED98 2824404 2009/11/19 2752<br />

S. aureus, a Commensal and Opportunistic Pathogen<br />

Staphylococcus aureus is a persistent commensal in the anterior nares <strong>of</strong> approximately 20 %<br />

<strong>of</strong> the human population (persistent carriers). The bacterium is found intermittently in 30 % to<br />

60 % <strong>of</strong> the population, and in non-carriers, which is the remaining fraction <strong>of</strong> the population,<br />

nasal swabs are never positive for S. aureus (Kluytmans et al. 1997, Wertheim et al. 2005). Some<br />

authors subdivide the group <strong>of</strong> intermittent carriers into intermittent and occasional carriers<br />

(Eriksen et al. 1995). More recent publications classify the carriage status only into two groups:<br />

persistent carriers and others (van Belkum et al. 2009).<br />

The occurrence <strong>of</strong> the two contrary groups <strong>of</strong> persistent and non-carriers strongly hints for an<br />

influence <strong>of</strong> the <strong>host</strong>’s immune system on the possibility for S. aureus to colonize persistently<br />

21


Maren Depke<br />

Introduction<br />

(Foster 2009). Others have already published polymorphisms associated with each phenotype,<br />

linking variants enhancing the immune response with reduced colonization risk (van den Akker et<br />

al. 2006, Emonts et al. 2007; Foster 2009). In the main site <strong>of</strong> colonization, the anterior nares,<br />

S. aureus uses its adhesins like ClfB, IsdA, SdrC, SdrD, and SasG to attach to the <strong>host</strong> tissue (Foster<br />

2009).<br />

Normally, such carriage does not cause any symptoms <strong>of</strong> illness. On the other hand, S. aureus<br />

can be responsible for a broad variety <strong>of</strong> diseases: S. aureus can cause mild to severe local<br />

infections <strong>of</strong> skin or pharyngeal mucosa, but also <strong>of</strong> inner organs (e. g. endocarditis,<br />

osteomyelitis) and systemic disease like sepsis. The <strong>pathogen</strong>icity <strong>of</strong> S. aureus is a world-wide<br />

problem. By the “SENTRY Surveillance Program”, S. aureus was described to be leading cause <strong>of</strong><br />

bloodstream, lower respiratory tract and skin/s<strong>of</strong>t tissues infections (Diekema et al. 2001).<br />

S. aureus produces several toxins that are responsible for toxin related diseases like toxic shock<br />

syndrome or scalded skin syndrome. Staphylococcal endotoxin can cause food poisoning in case<br />

<strong>of</strong> consumption <strong>of</strong> contaminated meals. S. aureus carriage is a risk factor for bacteremia. In the<br />

majority <strong>of</strong> bacteremia cases, the strain isolated from blood is identical with the nasal colonizing<br />

strain. But paradoxically, carriers have a lower mortality than non-carriers in case <strong>of</strong> bacteremia<br />

(van Eiff et al. 2001, Wertheim et al. 2004).<br />

S. aureus has been known to be an extracellular <strong>pathogen</strong> for long time (Finlay/Cossart 1997).<br />

But it has been also proven that S. aureus is not only an extracellular <strong>pathogen</strong> but that it can be<br />

internalized <strong>by</strong> non-pr<strong>of</strong>essional phagocytes like epithelial cells (Almeida et al. 1996, Hudson et<br />

al. 1995). Staphylococcal fibronectin-binding proteins (FnBP) are important mediators for the<br />

uptake <strong>of</strong> bacteria <strong>by</strong> the <strong>host</strong> cell, but it is not clear whether the uptake is an attribute <strong>of</strong><br />

bacterial <strong>pathogen</strong>icity or <strong>of</strong> <strong>host</strong> defense (Sinha et al. 2000). The process <strong>of</strong> internalization<br />

implies an active participation <strong>of</strong> <strong>host</strong> cells (Hudson et al. 1995).<br />

S. aureus Virulence Factors<br />

S. aureus has the genetic information to produce a wide range <strong>of</strong> virulence factors, which help<br />

to improve the bacterium’s survival e. g. in infection settings and to fight against the defense <strong>of</strong><br />

its <strong>host</strong>. Some are even required for its ability to establish an infection (Fig. I.4).<br />

S. aureus secretes several enzymes and exotoxins. A direct mode <strong>of</strong> impairing the <strong>host</strong>’s<br />

defense is to lyse <strong>host</strong> cells, which additionally renders nutrients accessible for the <strong>pathogen</strong>. As<br />

first virulence factors secreted enzymes (proteases, lipases, elastases, hyaluronidase and others)<br />

degrade <strong>host</strong> tissue molecules and therefore disintegrate tissue structure and defense barriers<br />

and facilitate first invasion and later spread <strong>of</strong> bacteria (Gordon RJ/Lowy 2008). All bacteria need<br />

iron for their survival. This essential nutrient is not freely available in the <strong>host</strong> but is sequestered<br />

<strong>by</strong> <strong>host</strong> proteins. S. aureus owns iron uptake promoting proteins, which are encoded <strong>by</strong> the ironresponsive<br />

surface determinant locus isd. This system includes binding proteins for different ironcontaining<br />

<strong>host</strong> proteins, transfer and transporter proteins, and proteins which release the iron<br />

into the bacterial metabolism (Maresso/Schneewind 2006).<br />

S. aureus carries different hemolysins which not only act on erythrocyte membranes, but also<br />

on that <strong>of</strong> other cell types. Alpha-hemolysin, which is also named alpha-toxin (encoded <strong>by</strong> the<br />

gene hla), is one <strong>of</strong> the prototypes <strong>of</strong> pore-forming toxins. This toxin has surprising characteristics<br />

<strong>of</strong> being water-soluble, but also able to form pores in hydrophobic membrane surrounding, and it<br />

22


Maren Depke<br />

Introduction<br />

contains all properties for its own insertion into membranes and therefore needs no assistance<br />

<strong>by</strong> proteins like chaperones (Menestrina et al. 2001). Other cytolytic toxins <strong>of</strong> S. aureus are betahemolysin<br />

(hlb), gamma-hemolysin (hlgA, hlgB, hlgC), delta-hemolysin (hld), and the groups <strong>of</strong><br />

leukocidins (luk). The Panton-Valentine leukocidin (PVL, lukF-PV and lukS-PV), a long-known and<br />

well-studied bicomponent toxin, depends in its pore-forming activity on two components, which<br />

first form themselves a heterodimer (Kaneko/Kamio 2004). Also the other leukocidins and Hlg are<br />

bicomponent lysins. Hlg can form either the AB toxin or the CB toxin depending on the<br />

association <strong>of</strong> the subunits to heterodimers.<br />

Fig. I.4: Staphylococcal structural and secreted virulence factors.<br />

A. Overview on surface and secreted proteins. TSST-1 – toxic shock syndrome toxin 1.<br />

B. Cell envelope structure and localization <strong>of</strong> selected proteins.<br />

C. Domains <strong>of</strong> cell wall anchored proteins.<br />

From: Gordon/Lowy 2008.<br />

Other important staphylococcal exotoxins are the so-called superantigens (SAg). These<br />

proteins disturb the <strong>host</strong>’s cellular adaptive immune response. Normally, T cells bind with their<br />

antigenic-peptide specific receptor (T-cell-receptor, TCR) to MHC-molecules, which present<br />

peptides derived from proteins inside or outside the antigen-presenting cell. This forms the socalled<br />

“immunological synapse” between antigen-presenting cell (APC) and T cell. In case <strong>of</strong> both<br />

specific recognition <strong>of</strong> the peptide <strong>by</strong> TCR as well as presence <strong>of</strong> further pro-stimulatory signals,<br />

the T cell will be activated, start to proliferate and secrete cytokines, and in general pursue their<br />

function in immune response. Superantigens interfere in this interaction. They can bind MHC-IImolecules<br />

on APC (with different preferences for MHC class II alleles) and also the TCR, in detail a<br />

subset <strong>of</strong> possible recombined TCR molecules characterized <strong>by</strong> certain Vβ-fragments in the β-<br />

chain (Herman et al. 1991). By this means, they build a bridge linking both receptors<br />

independently <strong>of</strong> an antigenic peptide and receptor specificity. This cross-linking in the<br />

trimolecular complex TCR/MHC-II/SAg induces a massive T cell proliferation and cytokine<br />

production leading to shock-like syndromes and generalized life-threatening damage. But the<br />

23


Maren Depke<br />

Introduction<br />

triggered <strong>host</strong> reaction does not help to fight the causative <strong>pathogen</strong>, because the T cell<br />

activation is antigen-independent and therefore unspecific to the <strong>pathogen</strong>. This, together with<br />

induction <strong>of</strong> immunosuppression, might be one <strong>of</strong> the advantages which superantigens grant the<br />

bacteria, although the question why bacteria harbor superantigens has not been completely<br />

clarified yet (Herman et al. 1991).<br />

Twenty superantigens <strong>of</strong> S. aureus have been described so far: toxic shock syndrome toxin-1<br />

(TSST-1), staphylococcal enterotoxins (SE A-E, G-J), and staphylococcal enterotoxin-like toxins (SEL<br />

K-R, U, U2, V). They cause diseases like toxic-shock syndrome and are linked to atopic dermatitis<br />

disease mechanism and allergic rhinitis (Fraser/Pr<strong>of</strong>t 2008). Mobile genetic elements <strong>of</strong> S. aureus<br />

encode the superantigens (Novick RP 2003).<br />

Further immunomodulation is achieved <strong>by</strong> extracellular adherence protein (Eap), which is also<br />

called MHC class II analogous protein (Map). This protein has similarity to MHC class II molecules<br />

(Jönsson et al. 1995) and influences T cell response during infection in favor <strong>of</strong> the infecting<br />

S. aureus (Lee et al. 2002). This inhibitory effect on T cells is dose-dependent and occurs only at<br />

high concentrations <strong>of</strong> Eap, while Eap stimulates T cell proliferation when present in low<br />

concentrations (Haggar et al. 2005).<br />

Virulence <strong>of</strong> S. aureus is promoted <strong>by</strong> the ability to attach to the <strong>host</strong>’s tissue, i. e. to the<br />

<strong>host</strong>’s extracellular matrix or the <strong>host</strong>’s cells, <strong>by</strong> its own surface components. Microbial surface<br />

components recognizing adhesive matrix molecules (MSCRAMMs) mediate the attachment,<br />

which can be regarded as the first initializing step for the establishment <strong>of</strong> infection<br />

(Gordon RJ/Lowy 2008). Molecules with similar function, but secreted in contrary to the<br />

microbial surface bound MSCRAMMs, are called secretable expanded repertoire adhesive<br />

molecules (SERAMs).<br />

Fibronectin binding proteins A and B (encoded <strong>by</strong> fnbA and fnbB) have binding properties for<br />

fibronectin, a high molecular weight protein <strong>of</strong> the <strong>host</strong>’s extracellular matrix (ECM). S. aureus<br />

also possesses adhesins binding to other ECM components like collagen (collagen binding protein<br />

Cna), elastin (elastin-binding protein EbpS). Clumping factor A and B (encoded <strong>by</strong> clfA and clfB)<br />

are fibrinogen binding proteins. This property results on the one hand in adhesin function <strong>of</strong><br />

clumping factor (Clf). But as Clf can not only bind to the blood coagulation precursor fibrinogen,<br />

but also activate fibrinogen to form fibrin, Clf has on the other hand agglutination properties.<br />

Structurally similar proteins to Clf exist. These are serine-aspartate (SD) repeat proteins, which<br />

are encoded <strong>by</strong> the genes sdrC, sdrD and sdrE, but their function is still unknown (Foster/Höök<br />

1998).<br />

A MSCRAMM-function <strong>of</strong> protein A is hypothesized from its ability to bind von-Willebrandfactor<br />

(vWF) in the environment <strong>of</strong> damaged epithelium where vWF originally enables the<br />

adhesion <strong>of</strong> platelets (Hartleib et al. 2000). Interestingly, a novel vWF binding protein (vWBP,<br />

encoded <strong>by</strong> vwb) <strong>of</strong> S. aureus has been discovered more recently (Bjerketorp et al. 2002).<br />

Fibronectin-binding protein (FnBP), clumping factor A (ClfA), protein A (Spa), and collagen<br />

binding protein (Cna) belong to the same group <strong>of</strong> MSCRAMMs. They contain a LPXTG motif<br />

which is cleaved during protein secretion through the cellular membrane with sequential<br />

anchoring <strong>of</strong> the remaining protein via the now accessible carboxylgroup <strong>of</strong> threonin (T) to the<br />

cell wall peptidoglycan (Foster/Höök 1998). Further MSCRAMMs in S. aureus are laminin-,<br />

vitronectin-, thrombospondin- and bone sialoprotein-binding proteins (Patti et al. 1994).<br />

MSCRAMMs do not only mediate the adhesion to the <strong>host</strong>’s extracellular matrix, but they are<br />

also involved in invasion into the <strong>host</strong> cells. Fibronectin-binding proteins attach the bacterial cell<br />

24


Maren Depke<br />

Introduction<br />

via fibronectin as linker to the <strong>host</strong>’s α 5 β 1 integrin molecules, which initiates the internalization <strong>of</strong><br />

the bacterial cell into the <strong>host</strong> cell (Hauck/Ohlsen 2006, Peacock et al. 1999, Schwarz-Linek et al.<br />

2004, Sinha et al. 1999). S. aureus strains harboring the SCCmec type I also possess the pls gene<br />

located therein. This gene encodes the plasmin sensitive surface protein Pls, which reduces,<br />

contrarily to the MSCRAMMs, the adherence <strong>of</strong> S. aureus to <strong>host</strong> structures and its cellular<br />

invasiveness. Very recently, it has been published that Pls causes this effect <strong>by</strong> steric hindrance<br />

and blocking adhesin and <strong>host</strong> factor interaction (Hussain et al. 2009).<br />

An example for a SERAM is the staphylococcal coagulase. The extracellular enzyme is able to<br />

bind prothrombin, which is the enzyme precursor for the coagulation activation <strong>by</strong> conversion <strong>of</strong><br />

fibrinogen to fibrin. Thrombin activation <strong>by</strong> staphylococcal coagulase does not occur <strong>by</strong><br />

proteolysis but <strong>by</strong> forming an active complex <strong>of</strong> both molecules in a stoichiometric ratio <strong>of</strong> 1:1,<br />

which is called staphylothrombin (Kawabata et al. 1985, Kawabata/Iwanaga 1994). Moreover,<br />

coagulase molecules can be attached to the staphylococcal cell surface where it can bind directly<br />

to fibrinogen also without presence <strong>of</strong> prothrombin (Bodén/Flock 1989).<br />

S. aureus is aiming to evade the immune response, and thus, it has gained several immunemodulating<br />

properties. After the very first steps <strong>of</strong> infection, the immune response has to be<br />

initiated <strong>by</strong> chemoattractants which recruit defense cells like neutrophils or macrophages to the<br />

site <strong>of</strong> infection. The bacterial protein chemotaxis inhibitory protein <strong>of</strong> staphylococci (CHIPS)<br />

blocks with two different domains the <strong>host</strong> cell receptors for the <strong>host</strong> chemoattractant C5a from<br />

the activated complement system and for bacterial formylated peptides, which are secreted <strong>by</strong><br />

the <strong>pathogen</strong> (de Haas et al. 2004, Murdoch/Finn 2000). Tightly associated with interference in<br />

the chemotaxic process is the already mentioned bacterial extracellular adherence protein (Eap).<br />

This protein blocks the endothelial cell receptor ICAM-1. During the normal immune response,<br />

ICAM-1 has receptor function for the leucocyte’s membrane protein LFA-1 and thus allows<br />

leucocyte adhesion at sites <strong>of</strong> infection. Adhesion is followed <strong>by</strong> entry <strong>of</strong> leucocytes into the<br />

tissue in the two steps <strong>of</strong> diapedesis and extravasation. In cases when ICAM-1 is already occupied<br />

<strong>by</strong> Eap, not only leucocyte adhesion but also the following steps are impeded (Chavakis et al.<br />

2002).<br />

S. aureus protects and hides its treacherous typical bacterial cell surface structures <strong>by</strong> a<br />

capsule <strong>of</strong> polysaccharides. The capsule inhibits binding <strong>of</strong> opsonic <strong>host</strong> factors and thus reduces<br />

the phagocytosis <strong>of</strong> the bacterial cell. An increased production <strong>of</strong> capsular polysaccharides<br />

increases the virulence <strong>of</strong> S. aureus (Nilsson et al. 1997, Thakker et al. 1998).<br />

Besides their already mentioned function <strong>of</strong> MSCRAMMs, clumping factor and protein A also<br />

have anti-phagocytic functions. Clumping factor binds the <strong>host</strong> molecule fibrinogen which then<br />

serves as camouflage net on the bacterial cell surface. Protein A binds immunoglobulin G (IgG) in<br />

a way beneficial for the bacterium, but not for the <strong>host</strong>. While the normal antibody-binding with<br />

the antigen-specific F ab part will opsonize the bacterium for recognition <strong>by</strong> immune defense<br />

mechanisms, the binding <strong>of</strong> antibodies <strong>by</strong> protein A occurs via the F c part and thus prevents the<br />

mediation <strong>of</strong> opsonization signals (Uhlén et al. 1984).<br />

Another protein which interferes with opsonization for phagocytosis is Staphylococcus<br />

complement inhibitor (SCIN). SCIN binds to the C3-convertase complexes from the three<br />

complement activation pathways and inhibits C3b deposition on the <strong>pathogen</strong>’s surface.<br />

Therefore, the following steps <strong>of</strong> complement opsonization are impaired. Interestingly, the<br />

complement-inhibitory function <strong>of</strong> SCIN is specific for the human <strong>host</strong> (Rooijakkers et al. 2005b).<br />

In a similar function, complement factor C3 is the target <strong>of</strong> the highly conserved, staphylococcal<br />

25


Maren Depke<br />

Introduction<br />

extracellular fibrinogen-binding protein Efb. Efb most probably inhibits the binding <strong>of</strong> C3b, the<br />

bigger fragment <strong>of</strong> C3 after cleavage during complement activation, to the bacterial surface and<br />

therefore reduces opsonization. Efb is able to bind both C3 and fibrinogen at the same time,<br />

because the C3-binding region is located C-terminal and the fibrinogen-binding region N-terminal<br />

in the Efb protein (Lee et al. 2004a, 2004b). More recently, a homologous and 44 % identical<br />

protein to Efb was identified. The protein called “Efb homologous protein”, Ehp, features the<br />

same C3b-deposition inhibitory function as Efb does, but in an even more potent manner. An<br />

additional C3-binding domain is proposed to cause this stronger inhibitory effect (Hammel et al.<br />

2007).<br />

S. aureus does not only use its own proteins for immune-modulatory purposes, but also<br />

engages and manipulates <strong>host</strong> factors to perform a protective effect for the bacterium. Host<br />

plasminogen is the inactive precursor <strong>of</strong> the serine protease plasmin, which is the main enzyme<br />

involved in fibrinolysis, the degradation <strong>of</strong> fibrin clots after injury and blood coagulation. Bacterial<br />

staphylokinase binds plasminogen to the staphylococcal cell surface and mediates the activation<br />

<strong>of</strong> the precursor into active plasmin. The plasminogen activation mechanism is not mediated via<br />

cleavage <strong>by</strong> staphylokinase as the bacterial protein does not possess enzymatic activity. More<br />

precisely, staphylokinase binds plasminogen in a 1:1 stoichiometric ratio and renders the enzyme<br />

precursor <strong>by</strong> formation changes more susceptible for activation. Activation finally occurs <strong>by</strong> other<br />

already active proteases, e. g. trace amounts <strong>of</strong> plasmin, which are able to start a reinforcing<br />

activation cascade (Silence et al. 1995). In a simple mode, active plasmin will help the <strong>pathogen</strong><br />

to spread from site <strong>of</strong> infection, either <strong>by</strong> enzymatic fibrinolysis <strong>of</strong> blood coagulation clots or <strong>by</strong><br />

degradation <strong>of</strong> extracellular matrix molecules. More sophisticated, the activated plasmin<br />

protease molecules degrade IgG and C3b from the bacterial surface and thus reverse<br />

opsonization (Rooijakkers et al. 2005a).<br />

When S. aureus is finally phagocytosed it has to deal with reactive oxygen species (ROS).<br />

Superoxide dismutase enzymes and non-enzymatic superoxide dismutase Mn 2+ inactivate<br />

superoxide anion O 2 – . Deletion mutants in superoxide dismutase or manganese cation uptake<br />

systems are less virulent than wild type strains. When ROS react with protein residues to<br />

methionine sulfoxide, these residues are reduced <strong>by</strong> methionine sulphoxide reductases, which<br />

also have impact on in vivo virulence (Foster 2005, Horsburgh et al. 2002a, Karavolos et al. 2003,<br />

Singh/Moskovitz 2003).<br />

Modulation <strong>of</strong> complement, opsonization and phagocytosis is not the only mechanism<br />

S. aureus applies for immune evasion. It also developed resistance mechanisms against killing <strong>by</strong><br />

antimicrobial peptides, especially after phagocytosis, when the four groups <strong>of</strong> antimicrobial<br />

substances interfere with the bacterial integrity: small anionic peptides (e. g. dermicidin in airway<br />

surfactant), small cationic peptides (e. g. cathelicidins in neutrophils), cationic disulphide bond<br />

forming peptides (e. g. α- and β-defensins), and anionic and cationic peptide fragments derived<br />

from larger proteins (Foster, 2005). In this context, staphylokinase has a further function. It binds<br />

α-defensins, which are secreted <strong>by</strong> neutrophils, and inhibits most <strong>of</strong> the defensins’ bactericidal<br />

potency (Jin et al. 2004). Vice versa, defensins inhibit the staphylokinase’s ability <strong>of</strong> activating<br />

plasminogen. This effect can be relevant in clinical settings when recombinant staphylokinase<br />

(sakSTAR) is planned to be developed into a thrombolytic pharmaceutical and might be applied in<br />

inflammation paralleled disorders like vascular occlusive diseases (Bokarewa/Tarkowski 2004).<br />

S. aureus secretes different proteases. One <strong>of</strong> them, aureolysin, cleaves the bactericidal peptide<br />

26


Maren Depke<br />

Introduction<br />

cathelicidin LL-37 at a special site, which is not recognized <strong>by</strong> the V8 protease. After aureolysin<br />

cleavage, LL-37 loses its antistaphylococcal activity (Sieprawska-Lupa et al. 2004).<br />

S. aureus cell wall is subjected to modifications <strong>of</strong> the teichoic acid structure. Dlt proteins<br />

(encoded <strong>by</strong> the dltABCD operon) lead to D-alanine incorporation. The alanine residues are<br />

further esterified. Without these alanine esters the bacterial surface would carry a strongly<br />

negative net charge, which would attract positively charged cationic antimicrobial molecules.<br />

Therefore, Dlt decreases the vulnerability <strong>of</strong> the bacterial cell wall via reduction <strong>of</strong> molecule<br />

charge (Peschel et al. 1999). On the other hand, a stronger negative charge would reduce<br />

S. aureus’ capacity to adhere to polystyrene or glass. Dlt mutants are thus impaired in their<br />

capability <strong>of</strong> bi<strong>of</strong>ilm formation (Gross et al. 2001). Also the MprF protein neutralizes charges <strong>of</strong><br />

the cell wall, in this case <strong>by</strong> linking lysine to phosphatidylglycerol (Peschel et al. 2001).<br />

Bi<strong>of</strong>ilm formation, which is <strong>of</strong>ten associated with intra-vascular medical devices like catheters,<br />

is another option for S. aureus to protect the bacterial cells from the immune system and<br />

antimicrobials. The material <strong>of</strong> the medical devices will be covered with <strong>host</strong> (serum) molecules<br />

like fibrinogen or fibronectin soon after implantation. Therefore, it serves as an ideal attachment<br />

site for the various adhesins <strong>of</strong> staphylococci (Lowy 1998). In bi<strong>of</strong>ilms, S. aureus aggregates in a<br />

multi-cellular structure which already hinders contact between immune cells and substances with<br />

bacterial cells <strong>by</strong> physical means. Bacterial cells are surrounded <strong>by</strong> a polysaccharide matrix,<br />

composed <strong>of</strong> poly-N-acetylglucosamine (PNAG). The ica operon, which is present in almost all<br />

S. aureus strains although not all <strong>of</strong> them form bi<strong>of</strong>ilms in in vitro assays, encodes the proteins<br />

necessary for PNAG biosynthesis. Its expression is regulated <strong>by</strong> IcaR and TcaR and the<br />

environmental conditions, and SarA and its homologues influence the regulation. Quorum<br />

sensing mechanisms are important in coordination <strong>of</strong> the individual bacterial cell’s gene<br />

expression, which enables the establishment <strong>of</strong> the bi<strong>of</strong>ilm (Fitzpatrick et al. 2005, Grinholc et al.<br />

2007).<br />

For bacteria, the most effective way <strong>of</strong> evading the immune response is to hide inside the <strong>host</strong><br />

cell. Intracellularly, S. aureus can mask its presence <strong>by</strong> changing its phenotype into the so-called<br />

small colony variant (SCV). These variants are characterized <strong>by</strong> physiological adaptation like slow<br />

growth and reduced toxin production, which is caused <strong>by</strong> alterations <strong>of</strong> the electron transport<br />

chain. They contribute to persistence and recurrence <strong>of</strong> staphylococcal infections<br />

(McNamara/Proctor 2000, von Eiff et al. 2006)<br />

Not all staphylococcal strains harbor the information for all virulence factors in their genome,<br />

and they do not necessarily express all virulence factors which are encoded. Virulence factor<br />

expression is limited <strong>by</strong> strict regulation to avoid wasting <strong>of</strong> energy or nutrient resources.<br />

Furthermore, the long list <strong>of</strong> different virulence factors and mechanism underlines the fact that<br />

S. aureus virulence factors are both redundant and multiple in function, with several factors<br />

performing the same or similar function and also one factor harboring several functions<br />

(Gordon RJ/Lowy 2008).<br />

27


Maren Depke<br />

Introduction<br />

Mechanisms <strong>of</strong> S. aureus Adaptation to its Environment<br />

For its survival in changing environments, the bacterium needs an adaptive response and thus<br />

an effective regulation <strong>of</strong> all cellular processes including gene expression, protein synthesis, and<br />

turnover allowing adaptation to different conditions. Such regulation not only controls cell<br />

division and metabolism, but also systems for exploiting limited nutrients, adaptation <strong>of</strong> the<br />

composition <strong>of</strong> the cell wall as outer boundary, surface factors like adhesins, and systems<br />

rendering stress resistance and guaranteeing endurance in situations non-favorable for growth or<br />

survival. In bacteria, the regulation <strong>of</strong>ten affects the transcription <strong>of</strong> genes as starting point for<br />

changes in the following levels like protein synthesis. Two-component systems (TCS) mediate the<br />

sensing <strong>of</strong> environmental signals with the sensor component and the adequate reaction with the<br />

response regulator component. An example for such TCS is the agr system. Additionally,<br />

transcription factors like SarA modulate staphylococcal gene expression (Cheung et al. 2004).<br />

Another well-conserved system is the use <strong>of</strong> alternative sigma factors. Transcription in<br />

bacteria is only initiated when the catalytic core complex <strong>of</strong> RNA polymerase, formed <strong>of</strong> the five<br />

subunits α 2 ββ’ω, is associated with a sigma factor recognizing the promoter (-10/-35-region). A<br />

so-called house-keeping sigma factor (in S. aureus: σ A ; Deora/Misra 1995) maintains the baseline<br />

expression <strong>of</strong> a “standard” set <strong>of</strong> genes that is generally needed <strong>by</strong> the cell. Alternative sigmafactors<br />

are activated in specific situations, e. g. after heat shock or salt stress. They recognize a<br />

specific set <strong>of</strong> promoters <strong>of</strong> genes whose function is required to encounter the corresponding<br />

physiological conditions. This set <strong>of</strong> genes includes further transcriptional regulators which<br />

themselves positively or negatively influence the expression <strong>of</strong> genes depending on the need <strong>of</strong><br />

the bacterial cell. S. aureus possesses three alternative sigma factors: σ B , σ H , and − most recently<br />

discovered − σ S (Wu et al. 1996; Morikawa et al. 2003; Shaw et al. 2008). SigB is homologous to<br />

sigB <strong>of</strong> Bacillus subtilis which has been subjected to intensive studies. Despite the similarity <strong>of</strong><br />

sigB in B. subtilis and S. aureus, only half <strong>of</strong> the gene cluster coding for proteins belonging to the<br />

control cascade <strong>of</strong> SigB activation/inactivation (rsbU-rsbV-rsbW-sigB) is conserved in S. aureus. Of<br />

special importance is the gene rsbU, coding for the phosphatase that is the starting point for<br />

activation <strong>of</strong> the alternative sigma factor SigB. After dephosphorylation <strong>by</strong> RsbU the anti-antisigma<br />

factor RsbV becomes active and binds RsbW leading to liberation <strong>of</strong> SigB from its antisigma<br />

factor RsbW. Contrarily to B. subtilis where further rsb gene products are regulators <strong>of</strong><br />

RsbU activity, in S. aureus an increase in RsbU already leads to SigB activation (Senn et al. 2005).<br />

Deletion <strong>of</strong> sigB in S. aureus leads among other things to a loss <strong>of</strong> pigmentation, increased<br />

sedimentation rate and increased sensitivity to hydrogen peroxide in the stationary growth<br />

phase. Different effects are observed on the expression <strong>of</strong> genes and on the abundance <strong>of</strong><br />

proteins. On the one hand some proteins are missing in sigB deletion mutants (e. g. Asp23) that<br />

are directly regulated and transcribed <strong>by</strong> SigB. On the other hand, other proteins have a higher<br />

abundance in the sigB deletion mutants (e. g. lipase, thermonuclease). For such genes a negative<br />

regulation <strong>by</strong> SigB itself or a SigB controlled regulator is basis <strong>of</strong> the regulation observed (Kullik et<br />

al. 1998). Bisch<strong>of</strong>f et al. (2004) propose YabJ, SpoVG, SarA, SarS and ArlRS as regulators possibly<br />

responsible for the indirect SigB effects.<br />

The SigB regulon in S. aureus has some similarities to that <strong>of</strong> B. subtilis, but the main function<br />

to obtain a broad-range stress resistance to the triggering stimulus and in advance also to other<br />

stressors is not conserved in S. aureus. Energy and ethanol stress do not trigger a SigB response<br />

in S. aureus. Nevertheless, some stress conditions such as heat shock, MnCl 2 , NaCl and alkaline<br />

28


Maren Depke<br />

Introduction<br />

shock are also effective in activating SigB in S. aureus. The entry into stationary growth phase<br />

additionally leads to the activation <strong>of</strong> the SigB regulon. Only a minor fraction <strong>of</strong> B. subtilis SigB<br />

dependent genes is found as orthologue in S. aureus and vice versa. In an approach to<br />

functionally characterize the SigB regulon <strong>of</strong> S. aureus, physiological aspects like cell envelope<br />

composition, membrane transport processes, and intermediary metabolism emerged as affected<br />

(Pané-Farré et al. 2006).<br />

The SigB regulon contains several virulence associated genes like coa, fnbA, ssaA, clfA, hla,<br />

hlgABC, lip, nuc, and sak. In a general view, a pattern <strong>of</strong> induction <strong>of</strong> cell surface virulence factors<br />

and a repression <strong>of</strong> exoproteins and toxins <strong>by</strong> SigB is recognized. Therefore, SigB effects are<br />

inverse to the effects <strong>of</strong> the active agr system via RNAIII (Bisch<strong>of</strong>f et al. 2004).<br />

Taken together, S. aureus exhibits a complex control <strong>of</strong> its virulence factors that includes<br />

several interlaced pathways <strong>of</strong> central regulators like agr, sarA, and sigB which are interesting<br />

targets for knock-out mutants to be tested in infection models.<br />

Increasing Importance and Danger <strong>of</strong> S. aureus Infections<br />

S. aureus is one <strong>of</strong> the main causes <strong>of</strong> hospital-and community-acquired infections, and its<br />

prevalence and antibiotic resistance have been studied intensively (Diekema et al. 2001, Goto et<br />

al. 2009).<br />

S. aureus as <strong>pathogen</strong> causing infection could be treated for a long time with classical<br />

antibiotics. But the <strong>pathogen</strong> has a potent ability to develop or acquire antibiotic resistances.<br />

With today’s increasing cases <strong>of</strong> antibiotic-resistant staphylococcal strains the question whether<br />

the era <strong>of</strong> untreatable infections has arrived is raised. This accentuated question emphasizes the<br />

recently enhanced importance and danger <strong>of</strong> S. aureus and complementarily, the imparative <strong>of</strong><br />

strict infection control and the importance <strong>of</strong> discovering new antibiotics and vaccination targets<br />

and developing these agents for medical purposes (Livermore 2009).<br />

Staphylococcal strains which are not only resistant to long- and <strong>of</strong>ten-used antibiotics, but<br />

also to those kept as reserve emerged and increase in prevalence in many countries (Livermore<br />

2000, Fig. I.5). This took place for methicillin (methicillin-resistant S. aureus – MRSA; MRSA might<br />

also be used in the context <strong>of</strong> multi-resistant S. aureus) since the 1960s, only some years after<br />

start <strong>of</strong> clinical application, or more recently for vancomycin (vancomycin-resistant<br />

S. aureus − VRSA) after S. aureus strains have obtained the vanA operon from an Enterococcus<br />

spp. transposon on a conjugative plasmid (Péricon/Courvalin 2009). The Staphylococcal Cassette<br />

Chromosome mec (SCCmec), a genetic region which is situated on a mobile genomic island and is<br />

thought to originate from Staphylococcus sciuri (Wu et al. 2001), is responsible for resistance to<br />

methicillin and all other ß-lactam antibiotics. This region exists in form <strong>of</strong> several different types<br />

(I-VIII) and can be used along with protein A typing (spa), multilocus sequence typing (MLST), and<br />

pulse-field gel electrophoresis (PFGE) to distinguish different staphylococcal lineages on a<br />

molecular level (Deurenberg/Stobberingh 2008, Zhang et al. 2009). Staphylococcal resistance to<br />

methicillin is based on different mechanisms. The main form is the expression <strong>of</strong> PBP2a, a nonmethicillin-sensitive<br />

variant <strong>of</strong> the sensitive original penicillin-binding proteins (PBPs). PBPs, <strong>of</strong><br />

which S. aureus owns four types PBP1-4, have essential function in cell wall biosynthesis, where<br />

they are responsible for cross-linking <strong>of</strong> peptide chains in the peptidoglycan and additionally<br />

stretch the glycan part via their transglycosylase activity. Methicillin inhibits PBP’s peptide cross-<br />

29


Maren Depke<br />

Introduction<br />

linking. In the presence <strong>of</strong> methicillin, MRSA induce the expression <strong>of</strong> PBP2a from their mecA<br />

gene, and survive antibiotic treatment because PBP2a resumes peptide cross-linking instead <strong>of</strong><br />

the inhibited standard PBPs. Other resistance mechanisms including a methicillin-specific<br />

lactamase (methicillinase) and PBP2 mutations with reduced methicillin binding are discussed<br />

(Chambers 1997, Stapleton/Taylor 2002).<br />

First, hospitals were the primary setting <strong>of</strong> transmission <strong>of</strong> and infection with such strains,<br />

certainly also favored <strong>by</strong> the gathering <strong>of</strong> people who might be regarded as pre-disposed victims<br />

for infection because <strong>of</strong> pre-existing illnesses or poor health state which additionally have<br />

frequent injections and insertions <strong>of</strong> medical devices (Lindsay/Holden 2004).<br />

But nowadays, MRSA has left the hospital as its transmission site and so-called “communityassociated”<br />

MRSA arise. Existence <strong>of</strong> this new group has been published beginning from 1990 for<br />

different countries. Interestingly, hospital-acquired / healthcare-associated (HA) and communityassociated<br />

(CA) MRSA consist <strong>of</strong> distinct strains, although <strong>of</strong> course CA-MRSA can cause<br />

nosocomial infections when transmitted <strong>by</strong> a carrier into hospital.<br />

CA-MRSA strains have characteristics which distinguish them from HA-MRSA. They <strong>of</strong>ten<br />

colonize non-classical niches in the human body and not the nares, and the distribution <strong>of</strong> the<br />

Staphylococcal Cassette Chromosome mec (SCCmec) types is different between HA- and CA-<br />

MRSA. Additionally, in CA-MRSA a higher frequency <strong>of</strong> Panton-Valentine leukocidin (PVL)-positive<br />

strains is found, although the PVL virulence factor is not a universal marker for CA-MRSA.<br />

Although the prevalence <strong>of</strong> CA-MRSA in Europe seems to be low, the number is already<br />

increasing and additionally might be underestimated because <strong>of</strong> difficulties in monitoring<br />

(symptom-free non-nasal carriage). Such unrecognized and hidden reservoir <strong>of</strong> MRSA is<br />

challenging with respect to attempts in controlling staphylococcal infectious disease<br />

(Otter/French 2010).<br />

A B C<br />

Fig. I.5: Proportion <strong>of</strong> MRSA isolates in EARSS-participating counties in 1999 (A), 2001 (B), and 2008 (C).<br />

Graphics from European Antimicrobial Resistance Surveillance System (EARSS), a european wide network <strong>of</strong> national surveillance<br />

systems, providing reference data on antimicrobial resistance for public health purposes (http://www.rivm.nl/earss/).<br />

30


Maren Depke<br />

Introduction<br />

STUDIES OF HOST-PATHOGEN INTERACTIONS<br />

Different approaches can be applied to define even more and specific <strong>interactions</strong> between<br />

the <strong>host</strong> and the <strong>pathogen</strong>. Besides classical physiological and microbiological methods, the<br />

modern molecular technologies can considerably help to improve the understanding <strong>of</strong> the<br />

processes acting during encounter <strong>of</strong> <strong>host</strong> and <strong>pathogen</strong> provided that genome sequence<br />

information is available. On the first level, RNA expression pr<strong>of</strong>iling will generate an overview on<br />

the potential changes <strong>of</strong> physiology and metabolism. This approach has the advantage <strong>of</strong><br />

monitoring the whole repertoire <strong>of</strong> the organism using a single analysis method and only one<br />

small sample, because nucleic acid material can be easily amplified <strong>by</strong> laboratory methods.<br />

Whether such changes <strong>of</strong> the RNA messenger really will be substantiated <strong>by</strong> changes on protein<br />

level must be analyzed <strong>by</strong> global or specific proteomic approaches. Here, different methods<br />

address the diverse cellular fractions and can provide a comprehensive impression <strong>of</strong> the<br />

different cellular states referring to protein abundance, but also to protein modification and<br />

subcellular protein localization, which is not accessible with the transcriptomic analysis. A<br />

limitation is in some cases the amount <strong>of</strong> sample material and the bigger effort needed to<br />

perform proteome studies in comparison to transcriptome studies. Results <strong>of</strong> both studies<br />

complement one another. Hence, a combination <strong>of</strong> several different approaches in co-operation<br />

studies is strongly recommended.<br />

Model Systems for Studies <strong>of</strong> Host Reactions Potentially Influencing the<br />

Outcome <strong>of</strong> Infections<br />

Liver gene expression pattern in a mouse psychological stress model<br />

Psychological and physiological stressors can disturb neuroendocrine, immunological,<br />

behavioral, and metabolic functions (Harris et al. 1998, Leibowitz/Wortley 2004, Mizock 1995)<br />

and adaptive physiological processes aim to reconstitute a dynamic equilibrium (McEwen 2004,<br />

Viswanathan/Dhabhar 2005).<br />

In a murine model <strong>of</strong> severe, chronic psychological stress due to 4.5 days <strong>of</strong> intermittent<br />

combined acoustic and restraint stress BALB/c mice developed severe systemic<br />

immunosuppression, neuroendocrinological disturbances and depression-like behavior. Besides<br />

heightened anti-inflammatory cytokine bias, lymphocytopenia, T cell anergy, impaired phagocytic<br />

and oxidative burst responses, increased susceptibility to experimental infection with E. coli,<br />

spontaneous bacterial infiltrations <strong>of</strong> gut commensals into the lung, reduced clearance <strong>of</strong><br />

experimental infections in the long term, attenuation <strong>of</strong> a hyperinflammatory septic shock, and<br />

finally, behavioral and neuroendocrine alterations and a prominent stress-induced loss <strong>of</strong> body<br />

mass without significant changes <strong>of</strong> food and water intake during the observation period became<br />

detectable (Kiank et al. 2006, 2007a, 2007b, 2008).<br />

31


Maren Depke<br />

Introduction<br />

Several authors propose chronic stress to be a main feature in the <strong>pathogen</strong>esis <strong>of</strong> the<br />

metabolic syndrome, which is associated with obesity, type 2 diabetes/insulin resistance,<br />

dyslipidemia, and hypertension (Alberti et al. 2009, Kuo et al. 2007, Lambert et al. 2010). On the<br />

other hand, stressors like injury, infection, traumatic events, or prolonged sleep deprivation<br />

capably induce hypercatabolism and, therefore, may cause cachexia (Alberda et al. 2006,<br />

Koban/Swinson 2005, Morley et al. 2006, Vanhorebeek/Van den Berghe 2004, van Waardenburg<br />

et al. 2006, Wilmore 2000, Wray et al. 2002). Such metabolic-driven wasting may result from<br />

pain, depression, or anxiety, causing malabsorption and maldigestion or morphological and<br />

functional alterations <strong>of</strong> the gastrointestinal system and is typically seen during repeated<br />

inflammatory processes or during sepsis (Alberda et al. 2006, Hang et al. 2003, Hansen MB et al.<br />

1998, McGuinness et al. 1995, Morley et al. 2006, van Waardenburg et al. 2006, Wilmore 2000).<br />

In the literature, two main pathways leading to massive loss <strong>of</strong> body mass are discussed: the<br />

response to starvation and the hypermetabolic response. Starvation is induced <strong>by</strong> inadequate<br />

calorie intake and causes the use <strong>of</strong> the body’s own tissue. At first, carbohydrate stores are<br />

emptied for energy production. Secondarily, the body switches to usage <strong>of</strong> lipid and protein<br />

stores for gluconeogenesis. Finally, fatty acids are absorbed into the liver where ketone bodies<br />

are produced that, for example, neurons can use as an energy source to restore functional<br />

homeostasis. In a feedback loop, efferent neurons signal to the periphery so that<br />

gluconeogenesis is lowered, and protein breakdown is diminished that causes adaptation to<br />

starvation. The supply for basal energy production then comes from the calories <strong>of</strong> adipose<br />

stores (Alberda et al. 2006, Morley et al. 2006).<br />

In contrast, a hypermetabolic response that is <strong>of</strong>ten seen during critical illness is<br />

predominantly mediated <strong>by</strong> hormones and inflammatory mediators. As during the initial phase <strong>of</strong><br />

starvation, gluconeogenesis is accelerated <strong>by</strong> usage <strong>of</strong> lipids and amino acids. Protein breakdown<br />

is massively increased due to glucagon and glucocorticoid effects, and can be accelerated <strong>by</strong><br />

proinflammatory cytokines such as TNF. Amino acids, along with fatty acids and glycerol, are used<br />

for gluconeogenesis in the liver, causing hyperglycemia. This process is mainly mediated <strong>by</strong><br />

catecholamines and corticosteroids, and finally results in a rapid loss <strong>of</strong> lean body mass without<br />

sufficient metabolic adaptation to diminish tissue breakdown (Costelli et al. 1993, Goodman<br />

1991, Harris et al. 1998, Lang et al. 1984, McGuinness et al. 1999, Mizock 1995, Morley et al.<br />

2006, Wilmore 2000, Vanhorebeek/Van den Berghe 2004, van Waardenburg et al. 2006, Weekers<br />

et al. 2002, Wray et al. 2002).<br />

To investigate primary causes for the severe loss <strong>of</strong> body mass in chronically stressed mice,<br />

gene expression pr<strong>of</strong>iles <strong>of</strong> the liver, which plays the major role in metabolism, were analyzed.<br />

Based on documented changes <strong>of</strong> the expression pr<strong>of</strong>ile <strong>of</strong> hepatic genes involved in<br />

carbohydrate, lipid, and amino acid metabolism, a <strong>characterization</strong> <strong>of</strong> metabolic disturbances<br />

was started and biologically relevant alterations <strong>of</strong> glucose metabolism, dyslipidemia, and<br />

changes in amino acid turnover in repeatedly stressed BALB/c mice were found which revealed a<br />

chronic stress-induced hypermetabolic syndrome (Depke et al. 2008).<br />

The liver fulfills an important function as a gatekeeper between the intestinal tract and the<br />

general circulation (Dietrich et al. 2003). Thus, the influence <strong>of</strong> psychological stress on immune<br />

regulatory processes in the liver was additionally investigated in the gene expression data set.<br />

Veins drain the intestinal tract and then segment to secondary capillary beds in the liver<br />

where the removal and metabolism <strong>of</strong> absorbed nutrients or toxic substances takes place. Food<br />

antigens and products <strong>of</strong> commensal bacteria are abundantly detectable in the portal vein and<br />

32


Maren Depke<br />

Introduction<br />

are effectively cleared <strong>by</strong> Kupffer cells and sinusoidal endothelial cells (Limmer et al. 2000,<br />

Maemura et al. 2005, van Oosten et al. 2001).<br />

Physiologically, there is no detectable intrahepatic immune activation in response to the low<br />

amount <strong>of</strong> microbial agent and food antigens derived from the gut because <strong>of</strong> a tolerogenic<br />

environment (Limmer et al. 2000, Macpherson et al. 2002). However, when the antigenic<br />

challenge is increased, an inflammatory response with a recruitment <strong>of</strong> neutrophils,<br />

monocytes/macrophages, and lymphocytes may be induced (Wiegard et al. 2005). Then, an<br />

increased production <strong>of</strong> inflammatory mediators such as reactive oxygen species (ROS) may<br />

cause cell damage and loss <strong>of</strong> hepatocyte functions (Ott et al. 2007).<br />

To elucidate hepatic immune regulatory pathways that may contribute to chronic<br />

psychological stress-induced immune suppression, the hepatic gene expression pr<strong>of</strong>iling was<br />

used to analyze expression changes <strong>of</strong> immune response and cell survival genes in stressed and<br />

non-stressed mice (Depke et al. 2009).<br />

Gene expression pattern <strong>of</strong> bone-marrow derived macrophages after interferon-gamma<br />

activation<br />

In the innate immune system, macrophages, together with dendritic cells, hold a central<br />

position. They are main effectors <strong>of</strong> the clearance <strong>of</strong> infections <strong>by</strong> their sentinel and phagocytic<br />

function. Macrophages present phagocytosed antigen derived peptides on MHC-II to<br />

lymphocytes. By this function, macrophages take part in regulation <strong>of</strong> the adaptive immune<br />

response (Gordon S 2007). Phagocytosis is mediated <strong>by</strong> different receptors like complement<br />

receptors, mannose receptor or via the interaction between lipopolysaccharide (LPS), LPS-binding<br />

protein (LBP), CD14, and Toll-like receptor TLR (Janeway/Medzhitov 2002, Greenberg/Grinstein<br />

2002). Upon binding <strong>of</strong> ligands to TLR, macrophages secrete chemokines like CXCL8, CXCL10,<br />

CCL3, CCL4, and CCL5, which mediate chemotaxis <strong>of</strong> neutrophils, NK cells, and T cells (Luster<br />

2002). Additionally, pro-inflammatory cytokines like TNF-α and IL-1 secreted <strong>by</strong> macrophages<br />

further contribute to inflammation. Macrophages promote extracellular matrix degradation <strong>by</strong><br />

their matrix metalloproteinases, MMPs (Gibbs et al. 1999a, 1999b). They exhibit increased killing<br />

activity towards phagocytosed <strong>pathogen</strong>s during respiratory burst, which is among others<br />

mediated <strong>by</strong> toxic, reactive defense molecules like NO and O –<br />

2 produced <strong>by</strong> inducible NO<br />

synthase (iNOS) and NADPH oxidase, respectively (DeLeo et al. 1999, Iles/Forman 2002, Kantari<br />

et al. 2008, MacMicking et al. 1997, Mori/Gotoh 2004, Park JB 2003). This macrophage<br />

phenotype corresponds to the classical activation and initiates primarily a Th1-based immune<br />

response (Fig. I.6). Vice versa, IFN-γ as part <strong>of</strong> the Th1 cytokine pattern provokes such classical<br />

activation (Duffield 2003).<br />

After a phase <strong>of</strong> inflammatory activation and phagocytosis <strong>of</strong> apoptotic cells at the site <strong>of</strong><br />

inflammation, macrophages are able to switch their functional pr<strong>of</strong>ile now aiming to reduce<br />

inflammation <strong>by</strong> anti-inflammatory cytokines, stop matrix degradation and even assist healing <strong>by</strong><br />

production <strong>of</strong> fibronectin and tissue transglutaminase. A similar phenotyp with reduced<br />

intracellular killing <strong>of</strong> <strong>pathogen</strong>s is observed after activation <strong>by</strong> Th2 cytokines like IL-4 and TGF-β.<br />

Activation leading to this phenotype is called “alternative” (Duffield 2003). Macrophages can be<br />

activated in a third way, called type II activation. This activation needs ligand binding to Fcγ<br />

receptors in combination with an activating stimulus like that mediated <strong>by</strong> TLRs. Afterwards, the<br />

33


Maren Depke<br />

Introduction<br />

phenotype <strong>of</strong> macrophages changes from production <strong>of</strong> IL-12 to IL-10 production, whereas TNFα,<br />

IL-1, and IL-6 are still secreted (Mosser 2003). Type II activation initiates a Th2-based immune<br />

response. There is experimental evidence that the phenotype <strong>of</strong> macrophages is primarily<br />

determined <strong>by</strong> the first cytokine type to which they are subjected (Erwig et al. 1998).<br />

A naïve macrophage can react to activation stimuli. Nevertheless, when the macrophage<br />

received a priming signal in advance, it had the possibility to prepare for the potentially following<br />

second stimulus <strong>by</strong> transcriptional regulation and is able to react stronger and faster to<br />

activation. Low-dose IFN-γ is the most important macrophage priming signal. Primed<br />

macrophages are not activated yet and do not secrete proinflammatory cytokines. This occurs in<br />

classical activation after a second signal like IFN-γ or LPS, peptidoglycan, and others (Dalton et al.<br />

1993, Huang et al. 1993, Ma J et al. 2003, Mosser 2003).<br />

Fig. I.6: Inducers and selected functional properties <strong>of</strong> different polarized macrophage populations.<br />

Abbreviations: DTH – delayed-type hypersensitivity; IC – immune complexes; IFN-γ – interferon-γ; iNOS – inducible nitric oxide<br />

synthase; LPS – lipopolysaccharide; MR – mannose receptor; PTX3 – the long pentraxin PTX3; RNI – reactive nitrogen intermediates;<br />

ROI – reactive oxygen intermediates; SLAM – signaling lymphocytic activation molecule; SRs – scavenger receptors; TLR – Toll-like<br />

receptor.<br />

From: Mantovani et al. 2004.<br />

Studies aiming to analyze the reaction in general and more specific the gene expression<br />

pattern <strong>of</strong> macrophages are impaired <strong>by</strong> the influence <strong>of</strong> immunological factors on the mature<br />

macrophages which can be prepared from animal organs, even under highly standardized<br />

laboratory conditions. This problem can be circumvented when using bone-marrow derived<br />

macrophages (BMM). Bone marrow stem cells are prepared from the animal and cultivated<br />

in vitro for proliferation and differentiation. Under the influence <strong>of</strong> granulocyte-macrophage<br />

colony stimulating factor (GM-CSF) cells differentiate into macrophages, which have the<br />

advantage <strong>of</strong> having never received any in vivo stimulus or immunological conditioning that<br />

might influence their cellular reaction.<br />

Until recently, further uncontrollable influences on macrophage or BMM experiments were<br />

introduced <strong>by</strong> the utilization <strong>of</strong> serum-supplemented culture medium. Serum contains immune<br />

34


Maren Depke<br />

Introduction<br />

cell stimulating substances like cytokines, hormones or endotoxin, and to worsen the problem,<br />

these substances might vary not only between the sera from different vendors but also between<br />

batches <strong>of</strong> serum from the same manufacturer. To overcome such sources <strong>of</strong> experimental<br />

variation, Eske et al. introduced in 2009 serum-free culture conditions for BMM. Standard cell<br />

culture medium RPMI 1640 was supplemented with a defined mixture <strong>of</strong> proteins, hormones,<br />

and other compounds. Differentiation <strong>of</strong> stem cells with recombinant murine GM-CSF yielded<br />

BMM expressing macrophage markers F4/80, CD11b, CD11c, MOMA-2, and CD13 after 10 days <strong>of</strong><br />

cultivation, and further characteristics <strong>of</strong> BMM differentiated in serum-containing medium were<br />

additionally retained (Eske et al. 2009).<br />

Host-<strong>pathogen</strong> interaction experiments have revealed differences in reactions <strong>of</strong> BMM<br />

derived from BALB/c and C57BL/6 mice when confronted with Burkholderia pseudomallei<br />

especially after IFN-γ stimulation and at higher multiplicities <strong>of</strong> infection (MOI). Also other<br />

infection studies uncovered differences between these two mouse strains in vivo and in vitro<br />

(Breitbach et al. 2006, Autenrieth et al. 1994, van Erp et al. 2006). The mouse strains BALB/c and<br />

C57BL/6 are characterized <strong>by</strong> a Th2 and Th1 centered type <strong>of</strong> immune response, respectively.<br />

Accordingly and possibly causatively, the main type <strong>of</strong> macrophage activation differs between the<br />

strains, with BALB/c preponderating the alternative macrophage activation pattern and C57BL/6<br />

prevailing classical activation (Mills et al. 2000).<br />

Against the background <strong>of</strong> genetic influences on the BMM reactions, a combined proteome<br />

(Dinh Hoang Dang Khoa) and transcriptome (Maren Depke) study was initiated. In a first<br />

experiment, BALB/c and C57BL/6 BMM were stimulated with IFN-γ to specify on a molecular level<br />

the reaction to the priming signal IFN-γ as basic principle. Furthermore, the study aimed to<br />

pr<strong>of</strong>ile potential differences <strong>of</strong> reaction between the BMM <strong>of</strong> both mouse strains.<br />

Model Systems for Studies <strong>of</strong> Host-Pathogen Interactions<br />

S. aureus strain RN1HG<br />

The genetic background <strong>of</strong> different S. aureus strains influences the reaction <strong>of</strong> the bacterium<br />

to experimental conditions. Therefore, the strain for experimental analysis must be chosen<br />

carefully because not all observations can be transferred from one to another S. aureus strain.<br />

The number <strong>of</strong> strains available for studies <strong>of</strong> S. aureus has grown recently when the group <strong>of</strong><br />

Friedrich Götz (Department <strong>of</strong> Microbial Genetics, Eberhards-Karls University <strong>of</strong> Tübingen,<br />

Germany) provided a rsbU + repaired RN1-derivative strain called RN1HG(001) and later tcaR + and<br />

rsbU + tcaR + repaired RN1-derivative strains to the scientific community (Herbert et al. 2010). The<br />

parental strain RN1 (NCTC8325) has already been widely used as model organism for diverse<br />

studies on staphylococcal physiology. On the other hand its defect in the important regulator<br />

rsbU was known (Kullik/Giachino 1997), which resulted in compromised conclusions from studies<br />

addressing the regulation <strong>of</strong> virulence factors (Giachino et al. 2001). Therefore, strain RN1HG in<br />

which the mutation in the regulatory gene rsbU has been complemented has been chosen in the<br />

experimental setup <strong>of</strong> the studies described in this thesis.<br />

35


infection rate<br />

cfu / organ<br />

Maren Depke<br />

Introduction<br />

Kidney gene expression pattern in an in vivo infection model<br />

S. aureus can be transmitted to the blood after body injury or <strong>by</strong> medical devices like<br />

catheters. An elementary model to mimic blood stream infection is the intra-venous infection <strong>of</strong><br />

laboratory animals, e. g. mice. Host reactions can be monitored <strong>by</strong> physiological, immunological<br />

or molecular readout systems. In this study, transcriptome analysis was applied.<br />

Strongest accumulation <strong>of</strong> S. aureus after i. v. infection <strong>of</strong> mice is observed in kidneys, which is<br />

also accompanied <strong>by</strong> bacterial proliferation in the time course <strong>of</strong> infection (Fig. I.7, data courtesy<br />

<strong>of</strong> Tina Schäfer). Thus, this organ was chosen for <strong>host</strong> gene expression pr<strong>of</strong>iling.<br />

Fig. I.7:<br />

Accumulation and bacterial proliferation <strong>of</strong> S. aureus<br />

Xen29 in murine kidney tissue after i. v. infection.<br />

Female BALB/c mice were infected with 1.0E+08 cfu<br />

via the tail vein. Data represent median and<br />

interquartile range <strong>of</strong> n = 3 experiments.<br />

Data courtesy <strong>of</strong> Tina Schäfer, Würzburg.<br />

1.0E+09<br />

1.0E+08<br />

1.0E+07<br />

1.0E+06<br />

1.0E+05<br />

1.0E+04<br />

1.0E+03<br />

1.0E+02<br />

1.0E+01<br />

0.25 24 48 72<br />

time post infection / h<br />

Although the virulence <strong>of</strong> sigB deficient strains is <strong>of</strong>ten reported to be similar to that <strong>of</strong> wild<br />

type strains the <strong>pathogen</strong>esis or pathomechanism <strong>of</strong> infection might be different. Therefore, the<br />

rationale <strong>of</strong> this study was to investigate whether the deletion <strong>of</strong> sigB will lead to a different<br />

reaction <strong>of</strong> the infected <strong>host</strong>.<br />

Host cell and <strong>pathogen</strong> gene expression pattern in an in vitro infection model<br />

While S. aureus colonizes humans in the anterior nares, it can be cause <strong>of</strong> pneumonia when<br />

transferred to the lung e. g. <strong>by</strong> aspiration or medical devices. In a longitudinal study with more<br />

than 10000 patients in Japan, S. aureus was identified as one <strong>of</strong> the leading causative organisms<br />

<strong>of</strong> pneumonia besides Streptococcus pneumonia and Haemophilus influenzae (Goto et al. 2009).<br />

Here, different cell types first encounter the <strong>pathogen</strong>: Cells associated with structural and<br />

functional aspects <strong>of</strong> lung and “guardian” cells <strong>of</strong> the immune system, e. g. alveolar macrophages.<br />

Epithelial cells form the inner surface <strong>of</strong> the lung. Cilial epithelial cells transport mucus out <strong>of</strong> the<br />

organ. The mucus is produced in the lower bronchia and alveoli <strong>by</strong> type II pneumocytes and<br />

functions as surfactant to reduce surface tension and as protective film to remove inhaled<br />

particles and <strong>pathogen</strong>s. Adjacent type I pneumocytes take part in oxygen exchange <strong>of</strong> the air<br />

with the blood.<br />

A model to study reactions <strong>of</strong> epithelial cells to <strong>pathogen</strong>s is the human bronchial epithelial<br />

cell line S9. S9 cells originate from the IB3-1 cell line, which was established about 20 years ago<br />

from a male, white cystic fibrosis (CF) patient’s bronchial epithelial cells after transformation with<br />

adeno-12-SV40 virus (Zeitlin et al. 1991). This cell line contains the non-functional chloride<br />

36


Maren Depke<br />

Introduction<br />

channel CFTR (cystic fibrosis transmembrane conductance regulator), which has been repaired in<br />

the S9 cell line <strong>by</strong> viral transfection with the wild-type gene (American Type Culture Collection<br />

ATCC, Manassas, VA, USA; www.atcc.org; S9 cell ATCC number CRL-2778).<br />

In vitro infection models can be accomplished <strong>by</strong> addition <strong>of</strong> three main types <strong>of</strong> bacterial<br />

supplement: 1) supernatant <strong>of</strong> bacterial cultures containing secreted proteins or virulence<br />

factors, 2) PBS-washed bacterial cells which bring only their membrane-bound and intracellular<br />

factors into the infection experiment, and 3) complete bacterial culture with both secreted<br />

proteins from supernatant and whole bacterial cells. This last option is nearest to the in vivo<br />

situation when the <strong>pathogen</strong> is able to influence its <strong>host</strong> with secreted factors. On the other<br />

hand, the established bacterial culture media, which are <strong>of</strong>ten lysates <strong>of</strong> protein-rich raw<br />

material (e. g. tryptic soy broth TSB), are highly artificial with reference to in vivo models or even<br />

to eukaryotic cell culture. Therefore, a medium was developed that allows bacterial growth but<br />

additionally has similarity to eukaryotic cell culture media (Schmidt et al. 2010). The authors<br />

describe the use <strong>of</strong> eukaryotic cell culture medium MEM supplemented with different amino<br />

acids, but without addition <strong>of</strong> serum. This new experimental system permits the study <strong>of</strong> <strong>host</strong><strong>pathogen</strong><br />

<strong>interactions</strong> in the context <strong>of</strong> all bacterial factors, membrane-bound and secreted, and<br />

additionally prevents effects on bacterial physiology <strong>by</strong> prolonged handling, centrifugation, and<br />

washing <strong>of</strong> bacteria.<br />

Here, exponential growth phase bacterial cultures were used to infect confluent S9 cell<br />

cultures. The strain S. aureus RN1HG has been chosen for a first insight into the molecular<br />

reactions in this model. RN1HG is a rsbU + repaired RN1-derivative strain (Herbert et al. 2010) with<br />

a SigB-positive phenotype.<br />

In a combined approach <strong>of</strong> transcriptome (Maren Depke) and proteome (Melanie Gutjahr)<br />

analysis the <strong>host</strong> reaction to infection and bacterial internalization was recorded. But not only<br />

stood the <strong>host</strong> cell in the focus <strong>of</strong> studies, but also the bacterium. In a similar experimental setup,<br />

internalized staphylococci were extracted from their S9 <strong>host</strong> cells and the bacterial RNA pr<strong>of</strong>ile<br />

was recorded using a tiling array approach (Maren Depke). Bacterial intracellular proteins were<br />

monitored and quantified after stable isotope labeling with amino acids in cell culture, SILAC<br />

(Sandra Scharf).<br />

Questions and Aims <strong>of</strong> the Studies Described in this Thesis<br />

This thesis contains results from transcriptome studies on different aspects <strong>of</strong> <strong>host</strong>-<strong>pathogen</strong><br />

<strong>interactions</strong>. First, liver gene expression pr<strong>of</strong>iles from a murine chronic stress model served to<br />

elucidate aspects <strong>of</strong> the influence <strong>of</strong> stress on the metabolism and the immune response state <strong>of</strong><br />

the animals. In the experiments for this study, the in vivo model <strong>of</strong> psychological stress in a<br />

complex mammalian <strong>host</strong> was performed without additional influences <strong>of</strong> a <strong>pathogen</strong>. Such<br />

influence was introduced in the second study: Here, the influence <strong>of</strong> staphylococcal i. v. infection<br />

on the <strong>host</strong> kidney gene expression was analyzed in another murine in vivo model using a wild<br />

type S. aureus strain and its isogenic sigB mutant. Tissue expression pr<strong>of</strong>iling from in vivo models<br />

has the advantage <strong>of</strong> directly recording the relevant physiological state with all its complex<br />

<strong>interactions</strong> and influences and its vicinity to medical questions in the human. Nevertheless, it is<br />

very difficult to distinguish the different components because the tissue samples are always a<br />

37


Maren Depke<br />

Introduction<br />

mixture <strong>of</strong> different cell types which might even feature contrary reactions. Therefore, in vitro<br />

models were additionally analyzed in which only one defined <strong>host</strong> cell type was studied.<br />

Macrophages are an example for an immune cell type involved in the first steps <strong>of</strong> the encounter<br />

between the <strong>host</strong> and a <strong>pathogen</strong>. A model to study reactions <strong>of</strong> macrophages is the preparation<br />

<strong>of</strong> bone marrow stem cells and the in vitro differentiation <strong>of</strong> the stem cells into so-called bonemarrow<br />

derived macrophages (BMM). The advantage <strong>of</strong> this approach is that these macrophages<br />

have never been under any immunological influence which might result from the immune status<br />

<strong>of</strong> the animal even under standardized laboratory conditions. Thus, the third part <strong>of</strong> this thesis<br />

focuses on the reaction <strong>of</strong> BMM. BMM <strong>of</strong> different mouse strains were treated with IFN-γ, a<br />

modulator <strong>of</strong> macrophage function which is one <strong>of</strong> the first signals during initiation <strong>of</strong> the<br />

immune response in vivo. But not only immune cells or specially phagocytes get in touch with<br />

<strong>pathogen</strong>s, but also cells responsible for functional and structural integrity <strong>of</strong> <strong>host</strong> organs and<br />

tissue, like epithelial and endothelial cells. Such cells are actually part <strong>of</strong> the first line <strong>of</strong><br />

recognition and reaction to a <strong>pathogen</strong>ic invasion into the <strong>host</strong>. The bronchial epithelial cell line<br />

S9 was used as an in vitro model system for the infection with staphylococci. The fourth chapter<br />

in this thesis includes <strong>host</strong> gene expression signatures <strong>of</strong> S9 cell after in vitro infection with<br />

S. aureus RN1HG. Finally, the following chapter addresses the <strong>pathogen</strong> expression pr<strong>of</strong>ile which<br />

was first recorded from agitated, aerobic S. aureus RN1HG cultures in different growth phases as<br />

a starting and reference point. Afterwards, the already described S9 cell in vitro infection model<br />

was used to extract staphylococci after an internalization phase inside the <strong>host</strong> cell. Internalized<br />

bacteria were analyzed at two time points in comparison to different control samples <strong>by</strong> tiling<br />

array gene expression analysis.<br />

38


Maren Depke<br />

M A T E R I A L A N D M E T H O D S<br />

LIVER GENE EXPRESSION PATTERN IN A MOUSE<br />

PSYCHOLOGICAL STRESS MODEL<br />

Animal experiments [performed <strong>by</strong> Cornelia Kiank]<br />

Female BALB/c mice aged 6 to 8 weeks were randomly grouped into the experimental and<br />

control groups starting at least 4 weeks before being used in experiments. The group size in<br />

different experiments differed from 6 to 12 mice per cage. Animals stayed in their group until the<br />

end <strong>of</strong> the experiments and were not mixed up to avoid social stress. All animals were<br />

maintained with sterilized food (ssniff R−Z; ssniff Spezialdiäten GmbH, Soest, Germany) and tap<br />

water ad libitum for adaptation under minimal stress conditions. Influences <strong>of</strong> irregularities <strong>of</strong><br />

the estrous cycles <strong>of</strong> unisexually grouped female mice were not analyzed selectively and may<br />

cause higher SD values in the statistical analyses.<br />

Animal rooms had a 12 h light, 12 h dark cycle and were maintained at a constant<br />

environment before the experiment. To avoid any additional effect, e. g. acoustic or olfactory<br />

effects, the handling <strong>of</strong> mice during the adaptation period and during the experiments was<br />

restricted to one investigator. All animal procedures were performed as approved <strong>by</strong> the Ethics<br />

Committee for Animal Care <strong>of</strong> Mecklenburg-Vorpommern, Germany.<br />

Repeated stress model [performed <strong>by</strong> Cornelia Kiank]<br />

Mice were exposed to combined acoustic and restraint stress on 4 successive days, for 2 h<br />

twice a day during the physiological recovery phase <strong>of</strong> rodents (0800–1000 and 1600–1800 h). On<br />

day 5, only one stress session was performed in the morning. For immobilization mice were<br />

placed in 50 ml conical centrifuge tubes with multiple ventilation holes without penning the tail.<br />

Acoustic stress was induced <strong>by</strong> a randomized ultrasound emission device between 19 kHz and<br />

25 kHz with 0 dB to 35 dB waves in attacks (patent no. 109977; SiXiS, Taipei, Taiwan), allowing<br />

the mice no adaptation to the stressor. Between the stress sessions, mice stayed in their home<br />

cages and had free access to food and tap water. Control mice were kept isolated from stressed<br />

animals during the 4.5 days stress exposure to avoid any acoustic or olfactory communication<br />

between the groups. Therefore, the nonstressed group stayed in the incubator where the<br />

animals were adapted. The stressed mice remained outside the incubator in the same animal<br />

laboratory during the whole period <strong>of</strong> the stress model. All successive experiments and analyses<br />

were performed starting at 1000 h after the ninth stress exposure. Different in vivo analyses were<br />

performed with 6 to 12 mice per group in at least two experiments according to the experimental<br />

protocol to ensure reproducibility. For array analysis two independent stress experiments were<br />

performed with nine mice per group (first experiment) and eight mice per group (second<br />

experiment).<br />

39


Maren Depke<br />

Material and Methods<br />

Liver Gene Expression Pattern in a Mouse Psychological Stress Model<br />

Acute stress model [performed <strong>by</strong> Cornelia Kiank]<br />

Mice were exposed to a single 2 h combined acoustic and restraint stress cycle in the morning<br />

(0800–1000 h). In vivo analyses were performed immediately after or 6 h after the stress session<br />

with six to nine mice per group. Two acute stress experiments for array analysis were composed<br />

<strong>of</strong> eight animals for each control group, and nine animals per stress group in the first and eight<br />

animals per stress group in the second experiment.<br />

Organ harvesting for RNA preparation [Cornelia Kiank / Maren Depke]<br />

Mice were killed <strong>by</strong> cervical dislocation, and organs were removed immediately to avoid RNA<br />

degradation. For liver samples, a small piece <strong>of</strong> tissue was immediately homogenized with a<br />

micropestle in 350 µl Buffer RLT / 1 % β-mercaptoethanol (QIAGEN GmbH, Hilden, Germany /<br />

Sigma-Aldrich Chemie GmbH, München, Germany). The liver lysates were shock frozen in liquid<br />

nitrogen. All samples were stored at −70°C.<br />

RNA preparation<br />

Liver sample lysates were thawed and processed at room temperature for RNA preparation<br />

with the RNeasy Mini Kit (QIAGEN GmbH, Hilden, Germany) according to the manufacturer’s<br />

instructions. After ethanol precipitation, the RNA was quantified spectrophotometrically, and its<br />

quality was verified using an Agilent 2100 Bioanalyzer and RNA Nano Chips (Agilent Technologies<br />

Inc., Santa Clara, CA, USA).<br />

DNA array analysis using Affymetrix expression arrays<br />

Pools containing equal amounts <strong>of</strong> RNA from each individual animal were prepared for each<br />

group and used for subsequent microarray analysis. Five micrograms <strong>of</strong> pooled total RNA were<br />

used for the synthesis <strong>of</strong> double-stranded cDNA, and this solution then served as a template for<br />

an in vitro-transcription reaction using GeneChip Expression 3’ Amplification Reagents<br />

(Affymetrix, Santa Clara, CA, USA) according to the manufacturer’s instructions. After spincolumn<br />

based cleanup, concentration and quality <strong>of</strong> cRNA were measured as described above.<br />

cRNA was fragmented, added to the hybridization cocktail, denatured, and hybridized with<br />

Affymetrix GeneChip Mouse Expression Arrays 430A/430A 2.0 according to the manufacturer’s<br />

instructions.<br />

Washing, staining, and scanning were performed using Affymetrix GeneChip FluidicsStation<br />

and scanners according to standard protocols.<br />

Data analysis for Affymetrix expression arrays<br />

The Affymetrix expression analysis was performed for the livers <strong>of</strong> repeatedly stressed and<br />

healthy control mice with technical duplicates <strong>of</strong> two independent biological experimental series<br />

each. For the analysis <strong>of</strong> the effects <strong>of</strong> acute stress, array hybridizations were also performed <strong>of</strong><br />

two independent biological experiments for both groups (control and acute stress). Affymetrix<br />

array image data generated with MAS 5.0 (repeated stress) were analyzed using the GeneChip<br />

Operating S<strong>of</strong>tware 1.2 (Affymetrix, Santa Clara, CA, USA) with default values for parameter<br />

settings. For normalization, a scaling procedure with a target value <strong>of</strong> 150 was used. Image data<br />

<strong>of</strong> the acute stress experiment were directly analyzed in GeneChip Operating S<strong>of</strong>tware 1.4 with<br />

default settings and normalized <strong>by</strong> scaling to the target value 500. After data transfer to the<br />

GeneSpring s<strong>of</strong>tware package (Agilent Technologies Inc., Santa Clara, CA, USA), genes displaying<br />

40


Maren Depke<br />

Material and Methods<br />

Liver Gene Expression Pattern in a Mouse Psychological Stress Model<br />

differential regulation in response to repeated and acute stress were identified based on the<br />

following criteria: 1) the signal <strong>of</strong> probe sets had to be present in the arrays at least in the control<br />

(for repressed genes) or in stressed mice (for induced genes) in both biological experiments,<br />

2) the difference <strong>of</strong> mean signals between control and stressed mice had to equal or exceed 100,<br />

and 3) the fold change factors calculated from the signal values in each experimental replicate<br />

had to exceed a cut<strong>of</strong>f <strong>of</strong> more than or equal to 1.5 or less than or equal to −1.5 in both biological<br />

experiments.<br />

Functional analysis <strong>of</strong> gene expression data using Ingenuity Pathway Analysis<br />

For data interpretation in the context <strong>of</strong> metabolism, lists <strong>of</strong> probe sets displaying differential<br />

regulation in both acute and repeated stress, or specifically after acute or repeated stress were<br />

uploaded as Excel spreadsheets (Micros<strong>of</strong>t Corp., Redmond, WA, USA) into the Ingenuity<br />

Pathway Analysis (IPA) Version 5.5 (Ingenuity Systems, Inc., Redwood City, CA, USA;<br />

http://www.ingenuity.com) and used for the interpretation <strong>of</strong> the array data in the context <strong>of</strong><br />

already published knowledge. Biological functions were assigned to the networks based on the<br />

content <strong>of</strong> the Ingenuity Pathway Knowledge Base. Complete hepatic gene expression data <strong>of</strong><br />

stressed and nonstressed mice are available at the NCBIs Gene Expression Omnibus (GEO,<br />

http://www.ncbi.nlm. nih.gov/geo/) database and are accessible through GEO Series accession<br />

no. GSE11126.<br />

For functional analysis in the context <strong>of</strong> immune response and apoptosis-associated genes,<br />

lists <strong>of</strong> differentially regulated probe sets were uploaded as Excel spreadsheets into the Ingenuity<br />

Pathway Analysis tool (IPA 6, www.ingenuity.com). In order to compare effects <strong>of</strong> acute and<br />

chronic stress, two groups <strong>of</strong> genes were included: (1) genes displaying differential regulation<br />

after acute stress and (2) genes differentially regulated after chronic stress. IPA combined the<br />

uploaded Affymetrix probe set IDs and assigned annotations depending on the content <strong>of</strong> the socalled<br />

Ingenuity Pathway Knowledge Base (IPKB). The IPKB was used to get further insight into<br />

the relation <strong>of</strong> differentially regulated genes and immunological functions. Searches for relevant<br />

keywords and meaningful combinations <strong>of</strong> keywords resulted in lists linked to the functions in<br />

focus. IPA also <strong>of</strong>fers the association <strong>of</strong> differentially expressed genes with canonical pathways,<br />

which can be rated <strong>by</strong> a corresponding p-value. This approach was used to depict functional<br />

aspects <strong>of</strong> differential gene expression.<br />

Real-time PCR<br />

DNA was removed <strong>by</strong> DNase treatment, and subsequent to purification using a RNeasy Micro<br />

Kit (QIAGEN GmbH, Hilden, Germany) and ethanol precipitation, concentration and quality <strong>of</strong><br />

RNA samples were assayed as described above. Validation <strong>of</strong> expression data <strong>by</strong> real-time PCR<br />

was separately performed for all individual RNA preparations (n = 9 plus n = 8 mice per group) <strong>of</strong><br />

the two biological experiments with repeated stress exposure. For real-time PCR analysis, 1 µg<br />

RNA was reverse transcribed into cDNA using the High Capacity cDNA Archive Kit in the presence<br />

<strong>of</strong> SUPERase•In RNase inhibitor (Ambion/Applied Biosystems, Foster City, CA, USA). Twenty<br />

nanograms <strong>of</strong> cDNA served as a template for real-time PCR using the following 20x TaqMan Gene<br />

Expression Assays (Applied Biosystems, Foster City, CA, USA): Asl (Mm00467107_m1), Srebf1<br />

(Mm00550338_m1), Pck1 (Mm00440636_m1), Gadd45b (Mm00435123_m1), Sds<br />

(Mm00455126_m1), and Actb (Mm00607939_s1). Differential regulation in repeatedly stressed<br />

and control mice was confirmed <strong>by</strong> comparing the ΔCt values (Ct value <strong>of</strong> the target<br />

gene − Ct value <strong>of</strong> the reference gene Actb in identical cDNA samples) <strong>of</strong> all control mice and<br />

repeatedly stressed mice with a Mann-Whitney U test, requiring a p-value <strong>of</strong> less than or equal to<br />

0.05.<br />

41


Maren Depke<br />

Material and Methods<br />

KIDNEY GENE EXPRESSION PATTERN IN AN<br />

IN VIVO INFECTION MODEL<br />

In vivo infection model, organ harvesting, and group size [performed <strong>by</strong> Tina Schäfer]<br />

Female BALB/c mice (Charles River, Sulzfeld, Germany) were infected i. v. with 5.0E+07 colony<br />

forming units (cfu; first biological replicate, BR1) and 7.0E+07 cfu (second biological replicate,<br />

BR2) S. aureus RN1HG or 5.0E+07 cfu (first biological replicate) and 8.0E+07 cfu (second biological<br />

replicate) <strong>of</strong> its isogenic sigB mutant S. aureus RN1HG ΔsigB. The third group in this study<br />

comprised sham-infected mice which received an injection <strong>of</strong> 100 µl physiological saline solution<br />

(performed only in the second biological replicate <strong>of</strong> the experiment). After 4 days (first biological<br />

replicate) or 5 days (second biological replicate) mice were sacrificed and kidneys were explanted<br />

immediately afterwards, flash-frozen in liquid nitrogen and stored at −80°C.<br />

Each <strong>of</strong> the three experimental groups was comprised <strong>of</strong> 5 independent samples per biological<br />

replicate (originating from kidneys <strong>of</strong> 5 mice) except for the group <strong>of</strong> infection with S. aureus<br />

RN1HG in the first biological replicate, which only consisted <strong>of</strong> 4 samples. In total, 24 samples<br />

were analyzed in this study.<br />

Tissue disruption<br />

In constant submersion in liquid nitrogen in a mortar, both kidneys <strong>of</strong> the mice were ground<br />

into very small pieces, but not into powder. By this means it was possible to yield a homogenous<br />

tissue mix which allowed accurate estimation <strong>of</strong> infection rate. As the tissue was not completely<br />

disrupted into powder, it was still possible to handle the frozen tissue mix e. g. during aliquoting<br />

without the risk <strong>of</strong> sample thawing.<br />

DNA preparation<br />

Small aliquots <strong>of</strong> disrupted tissue were added to Lysing Matrix D tubes (MP Biomedicals,<br />

Solon, OH, USA) which contain 1.4 mm diameter ceramic spheres. Tissue was completely<br />

disrupted in 190 µl <strong>of</strong> 42 mM EDTA using a FastPrep FP120 (Thermo Fisher Scientific Inc.,<br />

Waltham, MA, USA) at level 6.5 for 20 s. After short cooling on ice, samples were digested with<br />

250 µg/µl lysostaphin (AMBI PRODUCTS LLC, Lawrence, NY, USA) for 45 min at 37°C to ensure<br />

complete lysis <strong>of</strong> staphylococci in the infected tissue. The following DNA preparation employed<br />

the Wizard Genomic DNA Purification Kit (Promega Corp., Madison, WI, USA) with minor<br />

modifications to the manufacturer’s protocol. The tissue lysate (200 µl) was mixed with 830 µl <strong>of</strong><br />

Nuclei Lysis Solution (Promega) and incubated for 5 min at 80°C. Subsequently, samples were<br />

cooled for 5 min on ice and digested with 19.4 ng/µl RNase A (Promega) for 30 min at 37°C.<br />

Samples were again cooled on ice for 5 min and afterwards processed at room temperature.<br />

Lysates were transferred to new 1.5-ml-tubes, 345 µl <strong>of</strong> Protein Precipitation Solution were<br />

added, samples were vortexed for 20 s and protein was precipitated for 10 min on ice. The<br />

solution was cleared <strong>by</strong> two centrifugation steps at 20000 x g for 4 min (room temperature). The<br />

final clear supernatant was divided into two aliquots <strong>of</strong> 620 µl each and the DNA was precipitated<br />

<strong>by</strong> addition <strong>of</strong> 470 µl <strong>of</strong> isopropanol (2-propanol) and mixing <strong>by</strong> gentle inversion. After<br />

centrifugation for 2 min at 16000 x g (room temperature) the DNA pellet was washed with 600 µl<br />

43


Maren Depke<br />

Material and Methods<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

<strong>of</strong> 70 % ethanol, air dried for approximately 2 min and solved over night at 4°C in DNA<br />

Rehydration Solution (Promega, 10 mM Tris/HCl pH 7.4, 1 mM EDTA pH 8.0; 50 µl / pellet). Both<br />

aliquots <strong>of</strong> each sample were combined on the following day, the concentration was determined<br />

photometrically (NanoDrop ND-1000, NanoDrop Technologies, Wilmington, DE, USA), and the<br />

quality was checked <strong>by</strong> electrophoresis in 0.8 % agarose-TBE gels.<br />

Molecular estimation <strong>of</strong> infection rate with qPCR<br />

The DNA from infected tissue consists <strong>of</strong> a mix <strong>of</strong> <strong>host</strong> and <strong>pathogen</strong> DNA, which contains<br />

almost constant levels <strong>of</strong> <strong>host</strong> DNA and a varying proportion <strong>of</strong> bacterial DNA that very much<br />

depends on the infection rate. Therefore, quantifying the ratio <strong>of</strong> <strong>pathogen</strong> to <strong>host</strong> DNA allows<br />

an estimation <strong>of</strong> the infection rate, which can also be calibrated with the aid <strong>of</strong> artificially mixed<br />

standard samples.<br />

The highly conserved staphylococcal genes nuc (coding for thermonuclease) and dapA<br />

(encoding dihydrodipicolinate synthase) were quantified in reference to the mouse cytoskeleton<br />

gene Actb using 20x TaqMan Gene Expression Assays (Applied Biosystems, Foster City, CA, USA;<br />

Actb: inventoried assay Mm00607939_s1; nuc: custom assay, forward primer<br />

GATCCAACAGTATATAGTGCAACTTCAACT, reverse primer ACCGTATCACCATCAATCGCTTT, reporter<br />

AACCTGCGACATTAATT; dapA: custom assay, forward primer CTTTGAAGCGATTGCAGATGCT,<br />

reverse primer TGGTTCAATTGTCATGTTCGTTCTTG, reporter ACGACTGGTAATTTC; each reporter’s<br />

5’ end is labeled with FAM; a non-fluorescent quencher (NFQ) and a Minor Groove Binder (MGB)<br />

molecule is linked to the 3’ end).<br />

The real-time PCR was performed in triplicates with 20 mM Tris/HCl pH 7.4, 50 mM KCl, 3 mM<br />

MgCl 2 , 4 % glycerol, 0.2 mM dNTP mix (dATP, dCTP, dGTP, dTTP, 0.2 mM each), 1 % ROX reference<br />

dye (Invitrogen, Karlsruhe, Germany) and 0.35 U Platinum Taq DNA Polymerase (Invitrogen) using<br />

20 ng <strong>of</strong> DNA as template in each reaction. The so-called “universal settings” for Applied<br />

Biosystems’ TaqMan Assays were applied: 2 min at 50°C, 10 min at 95°C followed <strong>by</strong> 40 cycles <strong>of</strong><br />

15 s at 95°C and 1 min 60°C.<br />

A calibration curve with DNA prepared from a mixture <strong>of</strong> non-infected kidney tissue and<br />

in vitro cultivated staphylococcal cells was recorded in analogous reactions. Each standard<br />

contained a defined number <strong>of</strong> cfu per 10 mg tissue ranging from 1.32E+05 cfu/10 mg to<br />

3.31E+08 cfu/10 mg (Fig. M.2.1).<br />

A<br />

B<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

mean<br />

Linear (mean)<br />

y = -3.3265x + 23.246<br />

R² = 0.9946<br />

8.0<br />

6.0<br />

4.0<br />

2.0<br />

mean<br />

Linear (mean)<br />

y = -3.2883x + 22.211<br />

R² = 0.9956<br />

ΔCt<br />

0.0<br />

ΔCt 0.0<br />

-2.0<br />

-2.0<br />

-4.0<br />

-4.0<br />

-6.0<br />

-6.0<br />

-8.0<br />

5.0 6.0 7.0 8.0 9.0<br />

log ( cfu / 10 mg tissue)<br />

-8.0<br />

5.0 6.0 7.0 8.0 9.0<br />

log ( cfu / 10 mg tissue)<br />

Fig. M.2.1: Calibration curves for molecular estimation <strong>of</strong> infection rate using the staphylococcal genes dapA (A) and nuc (B).<br />

Mean values <strong>of</strong> ΔCt (ΔCt = Ct bacterial gene – Ct Actb) from n = 2 or n = 3 independent standard samples and the range (minimum to<br />

maximum) <strong>of</strong> these independent measurements are displayed.<br />

The Ct values <strong>of</strong> bacterial genes and Actb were used to calculate ΔCt<br />

(ΔCt = Ct bacterial gene − Ct Actb ). This value negatively correlates to the log-transformed values <strong>of</strong><br />

cfu/10 mg and can be described <strong>by</strong> a linear equation. The use <strong>of</strong> this linear correlation as<br />

calibration curve allows calculation the infection rate <strong>of</strong> the unknown experimental samples.<br />

44


Maren Depke<br />

Material and Methods<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

RNA preparation<br />

A frozen aliquot <strong>of</strong> disrupted tissue was disintegrated mechanically at 2600 rpm for 2 min in a<br />

bead mill (Mikrodismembrator S, B. Braun Biotech International GmbH, Melsungen, Germany;<br />

now part <strong>of</strong> Sartorius AG, Göttingen, Germany) with 0.5 ml TRIZOL (Invitrogen, Karlsruhe,<br />

Germany) using liquid nitrogen cooled Teflon vessels. Thereafter, another 0.5 ml TRIZOL was<br />

added to the still frozen lysate. After thawing, the lysate was incubated for 10 min at room<br />

temperature, flash-frozen in liquid nitrogen and stored −70°C until RNA preparation was<br />

continued. The lysates were again thawed at room temperature. Chlor<strong>of</strong>orm was added (200 µl<br />

chlor<strong>of</strong>orm / 1 ml TRIZOL), samples were shaken vigorously for 15 s and incubated at room<br />

temperature for 5 min. Organic and aqueous phase were separated <strong>by</strong> centrifugation (12000 x g,<br />

15 min, 4°C) and RNA was precipitated from the aqueous phase with 500 µl isopropanol (2-<br />

propanol) / 1 ml TRIZOL overnight at −20°C. After two washes with −20°C pre-cooled 80 % ethanol<br />

RNA was dried at room temperature and dissolved in nuclease-free water (Ambion Inc., Austin,<br />

TX, USA, now part <strong>of</strong> Applied Biosystems, Foster City, CA, USA).<br />

RNA was DNase treated and afterwards purified using the RNA Clean-Up and Concentration<br />

Kit (Norgen Biotek Corp., Thorold, ON, Canada; distributed <strong>by</strong> BioCat GmbH, Heidelberg,<br />

Germany). The concentration was determined photometrically (NanoDrop ND-1000, NanoDrop<br />

Technologies, Wilmington, DE, USA), and the quality was checked with an Agilent 2100<br />

Bioanalyzer (Agilent Technologies, Santa Clara, CA, USA).<br />

DNA array analysis<br />

The RNA expression pr<strong>of</strong>ile was analyzed on GeneChip Mouse Gene 1.0 ST arrays (Affymetrix,<br />

Santa Clara, CA, USA) using the Whole Transcript (WT) Sense Target Labeling and Control<br />

reagents according to the manufacturer’s instructions. Arrays were washed and stained in a<br />

GeneChip FluidicsStation 450 and scanned with a GeneChip Scanner 3000 (all: Affymetrix).<br />

DNA array data analysis<br />

The array image files (CEL) were first quality controlled in the Expression Console s<strong>of</strong>tware<br />

(Affymetrix) and then imported into the Rosetta Resolver s<strong>of</strong>tware (Rosetta Bios<strong>of</strong>tware, Seattle,<br />

WA, USA) for data analysis. Signals were generated and normalized using the RMA algorithm.<br />

Groups were compared in log-transformed space using error-weighted one-way ANOVA with<br />

Benjamini-Hochberg False Discovery Rate multiple testing correction and p* < 0.01 was regarded<br />

as significant. The control sequences (e. g. negative, positive, polyA, and hybridization controls)<br />

and sequences that were not expressed on all arrays in the selected group comparison (with<br />

p > 0.01 on intensity pr<strong>of</strong>ile level in Rosetta Resolver s<strong>of</strong>tware) were not included in statistical<br />

testing. Afterwards, sequence sets were translated into EntrezGene records using the Rosetta<br />

Resolver annotation <strong>of</strong> December 2009. Two criteria were used for definition <strong>of</strong> differential<br />

expression: Significance in ANOVA statistical testing and a minimal absolute fold change <strong>of</strong> 2.<br />

Genes significant in ANOVA but with a minimal absolute fold change <strong>of</strong> only 1.5 were considered<br />

to be regulated <strong>by</strong> trend.<br />

Ingenuity Pathway Analysis (IPA) <strong>of</strong> gene expression data<br />

EntrezGene identifiers and fold change gene expression data <strong>of</strong> differentially expressed genes<br />

were imported to the Ingenuity Pathway Analysis tool (Ingenuity Systems Inc.,<br />

www.ingenuity.com) and analyzed using all genes <strong>of</strong> the GeneChip Mouse Gene 1.0 ST array as<br />

reference set without further restriction.<br />

45


Maren Depke<br />

Material and Methods<br />

GENE EXPRESSION PATTERN OF BONE-MARROW DERIVED<br />

MACROPHAGES AFTER INTERFERON-GAMMA TREATMENT<br />

Stem cell preparation, cultivation, and differentiation to macrophages [performed <strong>by</strong><br />

Katrin Breitbach]<br />

Preparation and cultivation <strong>of</strong> mouse stem cells and their differentiation into bone-marrow<br />

derived macrophages (BMM) was conducted as described <strong>by</strong> Eske et al. in 2009. Briefly, bone<br />

marrow cells from tibias and femurs <strong>of</strong> n = 3 or n = 15 BALB/c or n = 4 or n = 15 C57BL/6 mice<br />

were prepared in sterile conditions, pooled and cultivated for ten days using the serum-free<br />

BMM-medium supplemented with GM-CSF as established <strong>by</strong> Eske et al. (2009). Differentiated<br />

BMM were harvested on day 10 for consecutive experiments.<br />

Interferon-γ activation <strong>of</strong> bone marrow derived macrophages [performed <strong>by</strong> Katrin Breitbach]<br />

Mature BMM were seeded in 6-well-plates with 0.8E+06 to 1.5E+06 cells / well. BMM in half <strong>of</strong><br />

the wells were treated for 24 hours <strong>by</strong> addition <strong>of</strong> 300 units/ml IFN-γ (Roche, Mannheim,<br />

Germany) in serum-free BMM-medium, the other half was cultivated for the same time in the<br />

same medium without IFN-γ. For transcriptome analysis, 1.6E+06 to 4.5E+06 cells / sample were<br />

available, while proteome analysis needed a higher cell number and therefore had a sample size<br />

<strong>of</strong> 1.0E+07 to 1.5E+07 cells.<br />

Sample [Katrin Breitbach] and RNA preparation [Maren Depke] for transcriptome analysis<br />

Medium was carefully removed from the sample wells, and BMM were lyzed in 1 ml<br />

TriReagent per sample (Sigma, Steinheim, Germany) <strong>by</strong> repetitive pipetting. After incubation at<br />

room temperature for 15 min, the lysate was flash-frozen in liquid nitrogen and stored at −70°C<br />

until RNA-preparation.<br />

RNA-preparation took place using a combined protocol <strong>of</strong> phenol-based preparation and<br />

column-based purification. Chlor<strong>of</strong>orm was added to TriReagent lysates, samples were vigorously<br />

shaken and incubated at room temperature for 5 min. Organic and aqueous phase were<br />

separated <strong>by</strong> centrifugation for 15 min at 12000 x g and 4°C. Afterwards, the aqueous<br />

supernatant was mixed with 0.5 ml isopropanol (2-propanol) and transferred to RNeasy Mini<br />

columns (Qiagen GmbH, Hilden, Germany). The following steps <strong>of</strong> RNA purification were carried<br />

out according to the manufacturer’s instructions including the optional DNase-treatment (RNasefree<br />

DNase Set, Qiagen GmbH, Hilden, Germany). After ethanol precipitation, the RNA was<br />

quantified spectrophotometrically, and its quality was verified using an Agilent 2100 Bioanalyzer<br />

and RNA Nano Chips (Agilent Technologies Inc., Santa Clara, CA, USA).<br />

Affymetrix DNA array analysis<br />

Four strain-treatment combinations were included into this study: 1) medium control BALB/c<br />

BMM, 2) IFN-γ treated BALB/c BMM, 3) medium control C57BL/6 BMM, and 4) IFN-γ treated<br />

C57BL/6 BMM. Three biological replicates were analyzed for each <strong>of</strong> the four sample groups<br />

described before.<br />

47


Maren Depke<br />

Material and Methods<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

Labeled cDNA for hybridization was prepared for each sample from 200 ng totalRNA using the<br />

GeneChip Whole Transcript (WT) Sense Target Labeling Assay and hybridized to GeneChip Mouse<br />

Gene 1.0 ST Arrays according to the manufacturer’s instructions (Affymetrix, Santa Clara, CA,<br />

USA). Arrays were washed and stained using the GeneChip Hybridization, Wash, and Stain Kit in a<br />

GeneChip Fluidics Station 450 and scanned with a GeneChip Scanner 3000 (all items from<br />

Affymetrix, Santa Clara, CA, USA).<br />

Resulting array image files (CEL-files) were imported to the Rosetta Resolver System (Rosetta<br />

Bios<strong>of</strong>tware, Seattle, WA, USA). Signal intensities were extracted using the RMA algorithm and<br />

differentially expressed probe sets were accessed with error-weighted one-way Analysis <strong>of</strong><br />

Variance (ANOVA) including Benjamini Hochberg False Discovery Rate (FDR) multiple testing<br />

correction at the analysis levels <strong>of</strong> Intensity Pr<strong>of</strong>iles and sequences in log-transformed space.<br />

Values <strong>of</strong> p* 0.01 on intensity pr<strong>of</strong>ile level in Rosetta Resolver s<strong>of</strong>tware) were not included in<br />

statistical testing.<br />

Each run <strong>of</strong> the statistical test compared two sample groups that consisted <strong>of</strong> three biological<br />

replicates each. Four comparisons were included into statistical testing: 1) IFN-γ treated BALB/c<br />

BMM versus medium control BALB/c BMM to access IFN-γ effects in BALB/c BMM, 2) IFN-γ<br />

treated C57BL/6 BMM versus medium control C57BL/6 BMM to retrieve IFN-γ effects in C57BL/6<br />

BMM, 3) medium control C57BL/6 BMM versus medium control BALB/c BMM to obtain strain<br />

differences at the non-activated control level, and 4) IFN-γ treated C57BL/6 BMM versus IFN-γ<br />

treated BALB/c BMM to elicit strain differences at the IFN-γ activated level.<br />

The Rosetta Resolver s<strong>of</strong>tware allows mapping <strong>of</strong> 28944 records <strong>of</strong> sequence level<br />

information (i.e. the Affymetrix probe sets) to genes via the EntrezGene nomenclature resulting<br />

in 20074 records. Additionally, the s<strong>of</strong>tware calculates expression data for genes from the<br />

original sequence level values. This function also combines intensities <strong>of</strong> two or more probe sets<br />

for genes that are represented <strong>by</strong> more than one probe set. The lists <strong>of</strong> statistically significant<br />

differentially expressed probe sets resulting from ANOVA were translated into lists <strong>of</strong><br />

differentially expressed genes <strong>by</strong> using the EntrezGene level annotation included in the Rosetta<br />

Resolver s<strong>of</strong>tware. In order to exclude biologically irrelevant small changes in expression level<br />

from the statistically significant lists resulting from ANOVA, expression data on EntrezGene level<br />

was restricted to a minimal absolute fold change <strong>of</strong> 1.5. Nevertheless, only a minority <strong>of</strong><br />

statistically significant genes did not pass this fold change cut<strong>of</strong>f.<br />

Proteome analysis [performed <strong>by</strong> Dinh Hoang Dang Khoa]<br />

Proteome analysis was performed <strong>by</strong> Dinh Hoang Dang Khoa using both gel-based and gelfree<br />

approaches. Briefly, BMM protein extracts in urea/thiourea buffer were separated <strong>by</strong> Two<br />

Dimensional Fluorescence Difference Gel Electrophoresis (2D-DIGE), and spots on the gels were<br />

identified using MALDI-MS-MS. Tryptic peptides <strong>of</strong> the protein extracts were analyzed with LTQ-<br />

FT-ICR mass spectrometer after liquid chromatographic (LC) prefractionation. Differential analysis<br />

<strong>of</strong> label-free MS data was performed using the Rosetta Elucidator s<strong>of</strong>tware package (Rosetta<br />

Bios<strong>of</strong>tware, Seattle, WA, USA).<br />

Comparison <strong>of</strong> transcriptome and gel-free LC-MS/MS proteome results<br />

To compare results from transcriptome and gelfree proteome analysis, it was necessary to<br />

map proteins from LC-MS/MS identification to genes available on the GeneChip Mouse Gene 1.0<br />

ST array (Affymetrix). In order to do so, protein IPI identifiers were translated into EntrezGene IDs<br />

using the PIPE (http://pipe.systemsbiology.net/pipe) and UniProt ID (www.uniprot.org) mapping<br />

tools (Lars Brinkmann). Missing EntrezGene IDs were added manually if possible. The resulting list<br />

48


Maren Depke<br />

Material and Methods<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

<strong>of</strong> proteins was imported into the Rosetta Resolver s<strong>of</strong>tware using the EntrezGene IDs and<br />

compared with genes on the Affymetrix array based on the annotation <strong>of</strong> the Rosetta Resolver<br />

s<strong>of</strong>tware.<br />

Global functional analysis <strong>of</strong> transcriptomic [Maren Depke] and proteomic [Dinh Hoang Dang<br />

Khoa] results using Ingenuity Pathway Analysis (IPA)<br />

The lists <strong>of</strong> differentially expressed genes with an absolute fold change <strong>of</strong> at least 1.5 were<br />

directly uploaded from Rosetta Resolver System to the Ingenuity Pathway Analysis tool Version<br />

7.5 and 8.0 (IPA, Ingenuity Systems, www.ingenuity.com). IPA assigned the EntrezGene IDs to the<br />

corresponding records <strong>of</strong> the so called Ingenuity Pathway Knowledge Base (IPKB), combined<br />

related genes to networks and conducted a statistical analysis for over-represented functions and<br />

canonical pathways in the imported lists <strong>of</strong> differentially expressed items in relation to the<br />

sequences available on the GeneChip Mouse Gene 1.0 ST Array as reference set. Generally,<br />

analysis in IPA was not restricted to species, type <strong>of</strong> relationship, type <strong>of</strong> molecule or the like.<br />

Only for building user-defined pathways, a restriction to cell type “macrophages” or “RAW cells”<br />

was used.<br />

Correspondingly, protein IPI identifiers, p-values, and fold change <strong>of</strong> regulated proteins were<br />

uploaded to IPA and also analyzed <strong>by</strong> the network, function and canonical pathway tools. In this<br />

case, the whole list <strong>of</strong> identified proteins was uploaded and afterwards restricted to a minimal<br />

absolute fold change <strong>of</strong> 1.5 and a p-value <strong>of</strong> at least 0.01. This approach allowed using all<br />

identified proteins as reference set.<br />

In case <strong>of</strong> interest, automatically generated networks from separate analyses were merged<br />

using the automatic merge-tool from IPA. User-defined networks (called pathways) centered<br />

around the starting node IFN-γ were created using the grow-tool <strong>of</strong> IPA‘s pathway function. The<br />

addition <strong>of</strong> further nodes according to information stored in the so-called Ingenuity Pathway<br />

Knowledge Base (IPKB) was restricted to a list <strong>of</strong> genes/proteins that are differentially expressed<br />

in at least 1 <strong>of</strong> 4 comparisons 1) IFN-γ effects in BALB/c BMM, 2) IFN-γ effects in C57BL/6 BMM,<br />

3) strain difference at non-treated control level and 4) strain difference after IFN-γ activation.<br />

Networks were built either without further restrictions or with additional restriction to<br />

macrophage/RAW cells (i.e. only genes/proteins in the IPKB described to be linked to IFN-γ in<br />

macrophages or RAW cells were allowed to enter the new network). Lists <strong>of</strong> genes included in<br />

the IFN-γ centered networks were compared to identify common and unique genes between the<br />

networks without and with restriction to macrophage/RAW cells and finally exported via<br />

spreadsheet.<br />

49


Maren Depke<br />

Material and Methods<br />

HOST CELL GENE EXPRESSION PATTERN IN AN<br />

IN VITRO INFECTION MODEL<br />

Host cell line and conditions <strong>of</strong> cell cultivation [performed <strong>by</strong> Melanie Gutjahr]<br />

S9 cells were grown to confluency in 10-cm-diameter cell culture dishes with 10 ml eMEM cell<br />

culture medium. Cell culture medium eMEM consists <strong>of</strong> 1x concentrated MEM (PromoCell GmbH,<br />

Heidelberg, Germany) supplemented with additional 4 % FCS (fetal calf serum, Biochrom AG,<br />

Berlin, Germany), 1 % non-essential amino acids (PAA Laboratories GmbH, Pasching, Austria), and<br />

2 % L-glutamine (200 mM stock; PAA).<br />

Bacterial growth medium and cultivation [performed <strong>by</strong> Melanie Gutjahr and Maren Depke]<br />

Staphylococcus aureus RN1HG GFP (plasmid pMV158GFP) was grown in 100 ml pMEM<br />

medium at 37°C with linear shaking <strong>of</strong> 125 strokes/min (stroke length 28 mm) in 500-ml-<br />

Erlenmeyer bacterial culture flasks. Optical density (OD) was measured at 600 nm.<br />

The adapted cell culture medium pMEM contains 1x concentrated MEM without sodium<br />

bicarbonate (10x concentrate; Invitrogen, Karlsruhe, Germany), 1 % non-essential amino acids<br />

(PAA Laboratories GmbH, Pasching, Austria) and 4 mM L-glutamine (PAA) and is supplemented<br />

with 10 mM HEPES (PAA) and 2 mM <strong>of</strong> each L-alanine, L-leucine, L-isoleucine, L-valine, L-<br />

aspartate, L-glutamate, L-serine, L-threonine, L-cysteine, L-proline, L-histidine, L-phenyl alanine,<br />

and L-tryptophan (PromoCell GmbH, Heidelberg, Germany). The pH is adjusted to 7.4 with NaOH.<br />

Use <strong>of</strong> this medium has been established <strong>by</strong> Sandra Scharf and has been first reported <strong>by</strong><br />

Schmidt et al. in 2010.<br />

Cell culture infection model [performed <strong>by</strong> Melanie Gutjahr and Maren Depke]<br />

When bacterial cultures reached an OD <strong>of</strong> 0.4 the S9 cells were infected with a multiplicity <strong>of</strong><br />

infection (MOI) <strong>of</strong> 25. Bacterial cultures and eukaryotic cell culture medium were mixed in a final<br />

volume sufficient for inoculation <strong>of</strong> all cell culture plates processed in parallel. Subsequently, this<br />

mix was added to the cell culture plates to ensure equal distribution <strong>of</strong> staphylococci on the <strong>host</strong><br />

cell layer and reproducible infection <strong>of</strong> all cell culture plates in each experiment.<br />

6.36E+07 cfu/ml had been determined at OD 0.4 <strong>of</strong> S. aureus RN1HG GFP in pMEM (Melanie<br />

Gutjahr). Confluent 10-cm-diameter cell culture plates contain 8.0E+06 S9 cells. Therefore, in<br />

these experiments approximately 30 % <strong>of</strong> bacterial culture had to be included in the infection<br />

medium to obtain a suspension <strong>of</strong> which 10 ml infect one cell culture plate with a MOI <strong>of</strong> 25<br />

(3.14 ml bacterial culture in a total <strong>of</strong> 10 ml infection medium).<br />

S9 cells and staphylococci were co-incubated for 1 h at 37°C in 5 % CO 2 -atmosphere.<br />

Afterwards, the infection medium was replaced <strong>by</strong> eMEM containing 10 µg/ml lysostaphin (AMBI<br />

PRODUCTS LLC, Lawrence, NY, USA) until harvesting <strong>of</strong> cells. Two time points were included in<br />

this study: 2.5 h and 6.5 h after start <strong>of</strong> infection (Fig. M.4.1).<br />

Control samples were treated accordingly. Only the volume <strong>of</strong> bacterial culture in the<br />

infection mix was substituted <strong>by</strong> fresh, sterile bacterial cell culture medium pMEM.<br />

51


Maren Depke<br />

Material and Methods<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

cultivation to OD 0.4<br />

experimental time point / h<br />

0 1 2 3 4 5<br />

6 7<br />

infection<br />

1h<br />

lysostaphin treatment<br />

t 0 2.5 h 6.5 h<br />

infected<br />

eukaryotic<br />

samples<br />

FACS for<br />

RNA and<br />

protein<br />

FACS for<br />

RNA and<br />

protein<br />

Fig. M.4.1:<br />

Time line <strong>of</strong><br />

infection<br />

experiments<br />

for eukaryotic<br />

<strong>host</strong> samples<br />

and control<br />

measurements.<br />

eukarotic<br />

control<br />

samples<br />

control<br />

measurements<br />

controlfor<br />

protein<br />

cfu<br />

<strong>host</strong> cell<br />

vitality<br />

controlfor<br />

RNA and<br />

protein<br />

cfu<br />

<strong>host</strong> cell<br />

vitality<br />

controlfor<br />

RNA and<br />

protein<br />

cfu<br />

<strong>host</strong> cell<br />

vitality<br />

Sample harvesting for transcriptome analysis<br />

After co-incubation <strong>of</strong> staphylococcal with <strong>host</strong> cells and subsequent lysostaphin treatment,<br />

the S9 cell layer was washed once with Dulbecco’s PBS without Ca 2+ and Mg 2+ (PAA Laboratories<br />

GmbH, Pasching, Austria). Cells were detached with 1 ml trypsin/EDTA (PAA) supplemented with<br />

0.1 µg/ml actinomycin D (Sigma-Aldrich, Steinheim, Germany) and 80 mM sodium azide (Merck<br />

KGaA, Darmstadt, Germany) for some minutes. Trypsin reaction was stopped with 3 ml cell<br />

culture medium eMEM, and cells were pelleted at room temperature for 5 min at 500 x g with<br />

slightly reduced break. The cells were washed once with Dulbecco’s PBS with Ca 2+ and Mg 2+ (PAA)<br />

and resuspended in FlacsFlow (Becton Dickinson Biosciences, San Jose, CA, USA) for sorting <strong>of</strong><br />

infected and non-infected <strong>host</strong> cells. Both solutions were supplemented with 0.025 µg/ml<br />

actinomycin D (Sigma-Aldrich) and 20 mM sodium azide (Merck). Control samples were treated in<br />

the same way.<br />

Protein samples and control measurements [performed <strong>by</strong> Melanie Gutjahr]<br />

Samples for transcriptome analysis were harvested in parallel with samples for <strong>host</strong> proteome<br />

analysis (Fig. M.4.1). For protein samples, trypsin, PBS and FacsFlow solutions were not<br />

supplemented with actinomycin D and sodium azide, and control samples were directly lysed in<br />

UT buffer (8 M urea, 2 M thiourea). One additional control without any treatment was included.<br />

For the bacterial starting culture (OD 0.4), the infection mix, and samples after 2.5 h and 6.5 h<br />

<strong>of</strong> infection (internalized bacteria) viable cell counts were determined <strong>by</strong> plating on TSB agar and<br />

incubation for 24-48 h at 37°C.<br />

FACS measurements and cell sorting [performed <strong>by</strong> Petra Hildebrandt], sample harvest and<br />

disruption [performed <strong>by</strong> Maren Depke]<br />

Cells were sorted into infected (green fluorescence positive) and non-infected (green<br />

fluorescence negative) cells in a biosafety 2 level FACS Aria high-speed cell sorter (Becton<br />

Dickinson Biosciences, San Jose, CA, USA) with 488 nm excitation from a blue Coherent Sapphire<br />

solid state laser at 18 mW. Optical filters were set up to detect the emitted GFP fluorescence at<br />

52


Maren Depke<br />

Material and Methods<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

515 nm to 545 nm (FITC filter block). All FSC and SSC data were recorded at linear scale, the<br />

fluorescence data were recorded at logarithmic scale with the FACS Diva v5.03 s<strong>of</strong>tware (Becton<br />

Dickinson). Prior to measurement <strong>of</strong> bacteria containing eukaryotic S9 cells, the proper function<br />

<strong>of</strong> the instrument was determined <strong>by</strong> using Spherotech’s Rainbow calibration particles<br />

(Spherotech, Lake Forest, IL, USA). Prior to sorting, the drop delay was adjusted to >99 % sorting<br />

with the ACCUDROP beads (Becton Dickinson). Sorting was performed from the gate set in FSC<br />

vs. FITC dot plot after eliminating doublets and cell clumps <strong>by</strong> gating SSC−W vs. SSC−A and FSC−W<br />

vs. FSC−A at a threshold rate up to 3000 cells/s with sort mode purity, which results in a sorted<br />

sample that is highly pure, at the expense <strong>of</strong> recovery and yield.<br />

For each transcriptome sample, 5.0E+05 to 7.0E+05 cells were collected, whereas for protein<br />

samples approximately 4.0E+06 cells were sorted. During the whole period <strong>of</strong> sorting and sample<br />

collection, the sample reservoir and the collection tubes were cooled to 4°C. Control samples<br />

were stored on ice during sorting <strong>of</strong> infected samples. Sorted and control cell suspensions were<br />

centrifuged for 5 min at 500 x g and 4°C. The supernatants were removed and each cell pellet<br />

devoted to transcriptome analysis was lysed in 1 ml TRIZOL (Invitrogen, Karlsruhe, Germany) <strong>by</strong><br />

repeated pipetting. After incubation for 10 min at room temperature, the lysate was flash-frozen<br />

in liquid nitrogen and stored −70°C until RNA preparation. The remaining cells after sorting <strong>of</strong><br />

samples for transcriptome and proteome analysis were used to acquire information on cell<br />

vitality <strong>by</strong> propidium iodide staining.<br />

RNA preparation<br />

The lysates were thawed at room temperature. Chlor<strong>of</strong>orm was added (200 µl<br />

chlor<strong>of</strong>orm / 1 ml TRIZOL), samples were shaken vigorously for 15 s and incubated at room<br />

temperature for 5 min. Organic and aqueous phase were separated <strong>by</strong> centrifugation (12000 x g,<br />

15 min, 4°C) and RNA was precipitated from the aqueous phase with 500 µl isopropanol (2-<br />

propanol) / 1 ml TRIZOL and 1 µl 5 mg/ml linear acrylamide (Ambion Inc., Austin, TX, USA, now<br />

part <strong>of</strong> Applied Biosystems, Foster City, CA, USA) as precipitation aid overnight at −20°C. After<br />

two washes with −20°C pre-cooled 80 % ethanol, the RNA was dried at room temperature and<br />

solved in nuclease-free water (Ambion). The concentration was determined photometrically<br />

(NanoDrop ND-1000, NanoDrop Technologies, Wilmington, DE, USA), and the quality was<br />

checked with an Agilent 2100 Bioanalyzer (Agilent Technologies, Santa Clara, CA, USA).<br />

DNA array analysis<br />

The RNA expression pr<strong>of</strong>ile <strong>of</strong> <strong>host</strong> cells was analyzed on GeneChip Human Gene 1.0 ST arrays<br />

(Affymetrix, Santa Clara, CA, USA) using the Whole Transcript (WT) Sense Target Labeling and<br />

Control reagents according to the manufacturer’s instructions. Arrays were washed and stained<br />

in a GeneChip FluidicsStation 450 and scanned with a GeneChip Scanner 3000 (all: Affymetrix).<br />

DNA array data analysis<br />

The array image files (CEL) were first quality controlled in the Expression Console s<strong>of</strong>tware<br />

(Affymetrix) and then imported into the Rosetta Resolver s<strong>of</strong>tware (Rosetta Bios<strong>of</strong>tware, Seattle,<br />

WA, USA) for data analysis. Signals were generated and normalized using the RMA algorithm.<br />

Groups were compared in log-transformed space using error-weighted one-way ANOVA with<br />

Benjamini-Hochberg False Discovery Rate multiple testing correction and p* < 0.01 was regarded<br />

as significant. The control sequences (e. g. negative, positive, polyA, and hybridization controls)<br />

and sequences that were not expressed on all arrays in the selected group comparison (with<br />

p > 0.01 on intensity pr<strong>of</strong>ile level in Rosetta Resolver s<strong>of</strong>tware) were not included in statistical<br />

testing.<br />

53


Maren Depke<br />

Material and Methods<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

Afterwards, the sequence sets were translated into EntrezGene records using the Rosetta<br />

Resolver annotation <strong>of</strong> December 2009, and an absolute fold change cut<strong>of</strong>f <strong>of</strong> 1.5 after<br />

combining the biological replicates was applied.<br />

Ingenuity Pathway Analysis (IPA) <strong>of</strong> gene expression data<br />

EntrezGene identifiers and fold change gene expression data <strong>of</strong> differentially expressed genes<br />

were imported to the Ingenuity Pathway Analysis tool version 8.6 (Ingenuity Systems Inc.,<br />

www.ingenuity.com) and analyzed using all genes <strong>of</strong> the GeneChip Human Gene 1.0 ST array as<br />

reference set without further restriction.<br />

54


Maren Depke<br />

Material and Methods<br />

PATHOGEN GENE EXPRESSION PROFILING<br />

Growth Media Comparison Study<br />

Bacterial cultivation, growth media, and sampling time points<br />

Staphylococcus aureus RN1HG was grown at 37°C with orbital shaking <strong>of</strong> 200 rpm in<br />

Erlenmeyer bacterial culture flasks. Culture volume did not exceed 1/5 <strong>of</strong> culture flask volume.<br />

Optical density (OD) was measured at 600 nm (Fig. M.5.1).<br />

Different media were included in this study in an international co-operation in the settings <strong>of</strong> the<br />

EU-IP-FP6-project BaSysBio (LSHG-CT2006-037469) consortium, e. g. the complete bacterial<br />

culture medium TSB, cell culture medium, minimal medium, and human serum. Here, only results<br />

from samples <strong>of</strong> the medium “pMEM” will be presented. The contents <strong>of</strong> the adapted cell culture<br />

medium pMEM (Schmidt et al. 2010) have already been listed above (see Material and<br />

Methods/Host Cell Gene Expression Pattern in an in vitro Infection Model/Bacterial growth<br />

medium, page 51).<br />

Bacterial samples were taken in the exponential growth phase and 2 h (t 2 ) and 4 h (t 4 ) after<br />

entry into stationary growth.<br />

OD600<br />

10.00<br />

1.00<br />

0.10<br />

TSB<br />

pMEM<br />

Fig. M.5.1:<br />

Example for bacterial growth in TSB and pMEM medium.<br />

0.01<br />

0 2 4 6 8 10 12 14 16 18 20 22 24<br />

time / h<br />

Bacterial cell harvest and disruption<br />

At different growth phases, 5 to 15 optical density units <strong>of</strong> S. aureus RN1HG were harvested<br />

on ice with addition <strong>of</strong> at least one third volume <strong>of</strong> Killing Buffer (20 mM Tris pH 7.5, 5 mM<br />

MgCl 2 , 20 mM NaN 3 ). Pellets were flash-frozen in liquid nitrogen and stored at −70°C until cell<br />

disruption.<br />

For cell disruption, the pellet was resuspended on ice in 200 µl Killing Buffer, transferred to<br />

liquid nitrogen cooled Teflon vessels, and disintegrated mechanically at 2600 rpm for 2 min in a<br />

bead mill (Mikrodismembrator S, B. Braun Biotech International GmbH, Melsungen, Germany;<br />

now part <strong>of</strong> Sartorius AG, Göttingen, Germany). Frozen cell and buffer powder mix was<br />

resuspendend in 4 ml <strong>of</strong> 50°C pre-warmed lysis solution (4 M guanidine-thiocyanate, 25 mM<br />

sodium acetate pH 5.5, 0.5 % N-lauroylsarcosinate) until the solution appeared clear and<br />

homogeneous. Four aliquots with 1 ml each were intermittently frozen in liquid nitrogen and<br />

stored −70°C.<br />

55


Maren Depke<br />

Material and Methods<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

RNA preparation<br />

RNA was prepared using the acid-phenol-method. The lysates / aqueous phases were<br />

extracted twice with an equal volume <strong>of</strong> acid phenol solution (Roth, Karlsruhe, Germany) and<br />

once with an equal volume <strong>of</strong> chlor<strong>of</strong>orm / isoamyl alcohol (3-methyl-1-butanol) mix (24 vol. <strong>of</strong><br />

chlor<strong>of</strong>orm and 1 vol. <strong>of</strong> isoamyl alcohol equilibrated with 1 M Tris/HCl pH 8.0). RNA was<br />

precipitated from the remaining cleaned aqueous phase with 1/10 volume <strong>of</strong> 3 M sodium acetate<br />

pH 5.5 and 1 volume <strong>of</strong> isopropanol (2-propanol) overnight at −20°C. After two washes with<br />

−20°C pre-cooled 80 % ethanol, the RNA was dried at room temperature and solved in nucleasefree<br />

water (Ambion Inc., Austin, TX, USA, now part <strong>of</strong> Applied Biosystems, Foster City, CA, USA).<br />

To avoid the influence <strong>of</strong> potential DNA contamination, the RNA was DNase treated and<br />

afterwards purified using the RNA Clean-Up and Concentration Kit (Norgen Biotek Corp., Thorold,<br />

ON, Canada; distributed <strong>by</strong> BioCat GmbH, Heidelberg, Germany). The concentration was<br />

determined photometrically (NanoDrop ND-1000, NanoDrop Technologies, Wilmington, DE,<br />

USA), and the quality was checked with an Agilent 2100 Bioanalyzer (Agilent Technologies, Santa<br />

Clara, CA, USA).<br />

56


Maren Depke<br />

Material and Methods<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

In vitro Infection Experiment Study<br />

Host cell line and culture conditions [performed <strong>by</strong> Petra Hildebrandt]<br />

S9 cells were cultivated as described above (see Material and Methods/Host Cell Gene<br />

Expression Pattern in an in vitro Infection Model, page 51).<br />

Bacterial growth medium and cultivation [performed <strong>by</strong> Melanie Gutjahr and Maren Depke]<br />

Staphylococcus aureus RN1HG or RN1HG GFP (plasmid pMV158GFP) was grown in 200 ml<br />

pMEM medium at 37°C with linear shaking <strong>of</strong> 110 strokes/min (stroke length 28 mm) in 1000-ml-<br />

Erlenmeyer bacterial culture flasks. All other experimental parameters (e. g. medium pMEM)<br />

were identical to those described above (see Material and Methods/Host Cell Gene Expression<br />

Pattern in an in vitro Infection Model, page 51).<br />

Cell culture infection model [performed <strong>by</strong> Melanie Gutjahr and Maren Depke]<br />

The cell culture infection model was performed as described above (see Material and<br />

Methods/Host Cell Gene Expression Pattern in an in vitro Infection Model, page 51) and also<br />

included the same two time points <strong>of</strong> 2.5 h and 6.5 h after start <strong>of</strong> infection (Fig. M.5.2).<br />

cultivation to OD 0.4<br />

experimental time point / h<br />

0 1 2 3 4 5<br />

6 7<br />

infection<br />

1h<br />

lysostaphin treatment<br />

t 0 1 h<br />

2.5 h 6.5 h<br />

internalized<br />

bacterial<br />

samples<br />

internalized<br />

S. aureus<br />

internalized<br />

S. aureus<br />

Fig. M.5.2:<br />

Time line<br />

<strong>of</strong> infection<br />

experiments<br />

for bacterial<br />

samples.<br />

bacterial<br />

control<br />

samples<br />

exponential<br />

growth<br />

phase<br />

(OD 0.4)<br />

S. aureus<br />

nonadherent<br />

S. aureus<br />

serum/CO 2<br />

control<br />

S. aureus<br />

anaerobic<br />

S. aureus<br />

serum/CO 2<br />

control<br />

S. aureus<br />

serum/CO 2<br />

control<br />

S. aureus<br />

57


Maren Depke<br />

Material and Methods<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

Bacterial control samples<br />

Bacterial samples were taken as comparison and to control for additional experimental effects<br />

(Fig. M.5.2 and Fig. M.5.3):<br />

exponential growth phase at OD 0.4<br />

staphylococci after 1 h, 2.5 h, and 6.5 h <strong>of</strong> incubation in the infection medium in 5 %<br />

CO 2 -atmosphere at 37°C without agitation (i.e. in cell culture dishes in serum<br />

containing medium but without presence <strong>of</strong> <strong>host</strong> cells)<br />

non-adherent staphylococci after 1 h <strong>of</strong> co-incubation with eukaryotic cells and their<br />

products in 5 % CO 2 -atmosphere and at 37°C<br />

staphylococci after 2.5 h <strong>of</strong> anaerobic incubation in pMEM medium at 37°C<br />

Cells were harvested using Killing Buffer (20 mM Tris pH 7.5, 5 mM MgCl 2 , 20 mM NaN 3 ) as<br />

described above (see Material and Methods/Growth Media Comparison Study/Bacterial cell<br />

harvest, page 55).<br />

exponential growth phase<br />

anaerobic<br />

incubation:<br />

2.5 h<br />

infection<br />

mix<br />

serum control<br />

with CO 2 exposure:<br />

1 h, 2.5 h, 6.5 h<br />

non-adherent staphylococci: 1 h<br />

Fig. M.5.3:<br />

Visualization <strong>of</strong> bacterial control samples in<br />

the in vitro infection experiments for tiling<br />

array analysis.<br />

internalized staphylococci: 2.5 h, 6.5 h<br />

S9 cells<br />

staphylococci<br />

Preparation <strong>of</strong> internalized staphylococci<br />

The lysostaphin-containing medium was replaced <strong>by</strong> 1 ml Killing Buffer (20 mM Tris pH 7.5,<br />

5 mM MgCl 2 , 20 mM NaN 3 ) with 150 mM NaCl. In this isotonic buffer the infected S9 cells were<br />

scraped from the culture dish, resuspended and transferred to a 1.5-ml tube for further<br />

processing while maintaining the integrity <strong>of</strong> the majority <strong>of</strong> <strong>host</strong> cells. All following steps were<br />

performed on ice or at 4°C. Cells were pelleted (600 x g for 5 min) and fixed in ice-cold<br />

acetone/ethanol (50 % v/v) for 4 min as described <strong>by</strong> Garzoni et al. (2007) followed <strong>by</strong> a<br />

centrifugation step at 20000 x g for 3 min. The eukaryotic part <strong>of</strong> the cell pellet was lysed in RLT<br />

buffer (Qiagen, Hilden, Germany) and homogenized twice using QIAshredder (Qiagen, Hilden,<br />

Germany) and centrifugation at 20000 x g for 2 min. In this process staphylococci were not lysed<br />

although they lost their viability. Therefore, the resulting pellet contained staphylococcal cells<br />

and eukaryotic cell debris. Pellets from several plates processed in parallel were combined and<br />

washed once with RLT buffer and four times with TE buffer (10 mM Tris/HCl pH 8, 1 mM EDTA<br />

pH 8) to remove residual contaminations <strong>of</strong> the <strong>host</strong> cells. Staphylococcal cell pellets were flashfrozen<br />

in liquid nitrogen and stored at −70°C until cell disruption.<br />

58


Maren Depke<br />

Material and Methods<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

Bacterial cell disruption<br />

For cell disruption the bacterial cell pellet was resuspended on ice in 60 µl TE (10 mM Tris/HCl<br />

pH 8, 1 mM EDTA pH 8) supplemented with 20 mM NaN 3 and transferred together with 0.5 ml<br />

TRIZOL (Invitrogen, Karlsruhe, Germany) to liquid nitrogen cooled Teflon vessels. Cells were<br />

disintegrated mechanically at 2600 rpm for 2 min in a bead mill (Mikrodismembrator S, B. Braun<br />

Biotech International GmbH, Melsungen, Germany; now part <strong>of</strong> Sartorius AG, Göttingen,<br />

Germany). Another 0.5 ml <strong>of</strong> TRIZOL was added and the frozen lysate was allowed to thaw. In<br />

total, the liquid lysate was incubated at room temperature for 10 min and afterwards flashfrozen<br />

in liquid nitrogen and stored −70°C until RNA preparation.<br />

RNA preparation<br />

The lysates from bead mill disruption were thawed at room temperature. Chlor<strong>of</strong>orm was<br />

added (200 µl chlor<strong>of</strong>orm / 1 ml TRIZOL), samples were shaken vigorously for 15 s and incubated<br />

at room temperature for 5 min. Organic and aqueous phase were separated <strong>by</strong> centrifugation<br />

(12000 x g, 15 min, 4°C) and RNA was precipitated from the aqueous phase with 500 µl<br />

isopropanol (2-propanol) / 1 ml TRIZOL overnight at −20°C. After two washes with −20°C precooled<br />

80 % ethanol the RNA was dried at room temperature and solved in nuclease-free water<br />

(Ambion Inc., Austin, TX, USA, now part <strong>of</strong> Applied Biosystems, Foster City, CA, USA).<br />

DNase treatment was only possible for control samples because the RNA yield <strong>of</strong> internalized<br />

staphylococci was near the minimum amount needed for tiling array hybridization. For controls,<br />

DNase digestion and RNA purification was performed as described before (see Material and<br />

Methods/Growth Media Comparison Study, page 56). The concentration was determined<br />

photometrically (NanoDrop ND-1000, NanoDrop Technologies, Wilmington, DE, USA) and the<br />

quality was checked with an Agilent 2100 Bioanalyzer (Agilent Technologies, Santa Clara, CA,<br />

USA).<br />

59


Maren Depke<br />

Material and Methods<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

Tiling Array Expression Pr<strong>of</strong>iling<br />

Tiling array design<br />

The “080604_SA_JH_Tiling” array was designed <strong>by</strong> Hanne Jarmer (Center for Biological<br />

Sequence Analysis, Department <strong>of</strong> Systems Biology, Technical University <strong>of</strong> Denmark, Lyng<strong>by</strong>,<br />

Denmark) using algorithms, which have recently been published in the context <strong>of</strong> a B. subtilis<br />

tiling array (Rasmussen et al. 2009). In brief, iso-thermal probes <strong>of</strong> 45 nucleotides (nt) to 65 nt<br />

were designed to cover the whole genome <strong>of</strong> S. aureus. Probes were arranged in 22 nt intervals<br />

on each strand and possessed an <strong>of</strong>fset <strong>of</strong> 11 nt between the strands. The array design for<br />

S. aureus was part <strong>of</strong> the international cooperation EU-IP-FP6-project BaSysBio (LSHG-CT2006-<br />

037469) which is funded <strong>by</strong> the European Commission and coordinated <strong>by</strong> Philippe Noirot (INRA,<br />

Mathématique Informatique et Génome, Jouy-en-Josas, France).<br />

The sequences were synthesized on quartz wafers <strong>by</strong> NimbleGen (Roche NimbleGen,<br />

Madison, WI, USA) in a custom, high-density DNA array format <strong>by</strong> applying the Maskless Array<br />

Synthesizer (MAS) technology in combination with photo-mediated synthesis chemistry. The<br />

basic principle <strong>of</strong> the synthesis is a solid-state array <strong>of</strong> miniature aluminum mirrors which direct<br />

UV-light at the place where the next reaction steps should occur. An UV-labile protection group is<br />

separated from the nascent oligonucleotide and releases a reactive site for binding <strong>of</strong> the next<br />

nucleotide. Thus, the synthesis <strong>of</strong> the oligonucleotide sequences requires m x 4 reaction steps,<br />

where m is the length <strong>of</strong> the oligonucleotide (number <strong>of</strong> nucleotides) and at each intermediate<br />

length step the four possible nucleotides A, T, C, and G will be added in a single, separate<br />

reaction.<br />

Tiling array hybridization<br />

The tiling array hybridization was performed at NimbleGen (Roche NimbleGen, Madison, WI,<br />

USA) according to standard protocols. Briefly, 10 µg <strong>of</strong> high quality RNA samples were reverse<br />

transcribed to cDNA in the presence <strong>of</strong> actinomycin D with subsequent alkaline RNA hydrolysis.<br />

Precipitated and resolved cDNA was labeled with Cy3 dye via NHS-Ester Dye Coupling Reaction.<br />

After spin-column based purification and quality control, labeled cDNA was hybridized together<br />

with appropriate controls to tiling arrays <strong>of</strong> the 080604_SA_JH_Tiling design. Arrays were washed<br />

and scanned and raw intensity data <strong>of</strong> tiling probes were provided to the customer.<br />

Tiling array raw data analysis<br />

The raw data analysis and condensing <strong>of</strong> probe data to transcripts/genes was performed <strong>by</strong><br />

Pierre Nicolas, Aurélie Leduc, and Philippe Bessières (INRA, Mathématique Informatique et<br />

Génome, Jouy-en-Josas, France). Intensity values for annotated genes were derived from the<br />

individual probe data <strong>by</strong> calculating the median <strong>of</strong> the probes located within the genomic<br />

coordinates <strong>of</strong> these genes. Analysis <strong>of</strong> hybridization signal, identification <strong>of</strong> transcription start<br />

and end boundaries and <strong>by</strong> this means segmentation into transcriptional units was executed <strong>by</strong> a<br />

novel algorithm based on a hidden Markov model, which allows to identify new, formerly<br />

unknown or non-annotated transcripts (Nicolas et al. 2009).<br />

60


Maren Depke<br />

Material and Methods<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

Tiling array expression data analysis<br />

Condensed, linear space gene expression values were imported via the Gene Expression Text<br />

Loader (GETL) into the Rosetta Resolver s<strong>of</strong>tware (Rosetta Bios<strong>of</strong>tware, Seattle, WA, USA) for<br />

data analysis. Inter-chip median-scaling (normalization) and detrending (Rosetta Resolver<br />

proprietary algorithm) were applied in the experiment definitions (ED) to allow direct comparison<br />

<strong>of</strong> values from different arrays. Groups were compared in log-transformed space using textbook<br />

one-way ANOVA with Benjamini-Hochberg False Discovery Rate multiple testing correction and<br />

p* < 0.05 was regarded as significant. An absolute fold change cut<strong>of</strong>f <strong>of</strong> 2 was applied.<br />

61


corticosterone [pg/ml plasma]<br />

corticosterone [pg/ml plasma]<br />

corticosterone [pg/ml plasma]<br />

Maren Depke<br />

R E S U L T S<br />

LIVER GENE EXPRESSION PATTERN IN A MOUSE<br />

PSYCHOLOGICAL STRESS MODEL<br />

The results <strong>of</strong> the liver gene expression pr<strong>of</strong>iling in a mouse psychological stress model have<br />

been published <strong>by</strong> Depke et al. in 2008 and 2009. All array data are available at the NCBI’s Gene<br />

Expression Omnibus (GEO, http://www.ncbi.nlm.nih.gov/geo/) database and are accessible<br />

through GEO Series accession no. GSE11126. Furthermore, lists <strong>of</strong> differentially expressed genes<br />

are included as supplemental data (Depke et al. 2008) at the journal’s homepage<br />

(http://endo.endojournals.org). Stress and infection experiments, physiological measurements,<br />

and histological analyses were performed <strong>by</strong> Cornelia Kiank. Her data are cited here because they<br />

are essential for the understanding <strong>of</strong> physiological events and potential dysregulations during<br />

acute and chronic stress in mice.<br />

Repeated stress-induced cachexia accompanied <strong>by</strong> hypercortisolism, hyperleptinemia, and<br />

hypothyroidism [Cornelia Kiank]<br />

Repeated psychological stress caused a severe loss <strong>of</strong> body mass in BALB/c mice while mean<br />

food intake and water consumption were unaltered during 4.5 days stress exposure. Food intake<br />

was 95.1 ± 19.9 g/cage in the repeatedly stressed vs. 91.9 ± 17.4 g/cage in the nonstressed<br />

groups, and water consumption was 250 ± 40 ml/cage in stressed vs. 240 ± 40 ml/cage in the<br />

control mice (three independent experiments with nine mice per cage). As the food and water<br />

intake was consistently found to be normal, the question arose whether the severe loss <strong>of</strong> body<br />

mass after repeated stress exposure was dependent on hormonal changes. Repeatedly stressed<br />

mice showed increased corticosterone concentrations in the peripheral blood (Fig. R.1.1 A) along<br />

with a hypertrophy <strong>of</strong> the adrenal cortex with decreased size <strong>of</strong> lipid storage vesicles in the<br />

glucocorticoid-producing zona fasciculata (Fig. R.1.1 B, C).<br />

A<br />

A<br />

A750 *<br />

A<br />

750 *<br />

750 *<br />

B<br />

B<br />

B<br />

B<br />

Fig. R.1.1: Repeated stress-induced activation <strong>of</strong> the HPA axis in BALB/c<br />

mice.<br />

A. Increased plasma corticosterone levels in repeatedly stressed mice<br />

(black box plot) compared with nonstressed mice (white box plot) (n = 9<br />

mice per group).<br />

B, C. Hypertrophy <strong>of</strong> the zona fasciculata <strong>of</strong> the adrenal cortex (white line)<br />

in repeatedly stressed mice (B) compared with nonstressed controls (C)<br />

(HE staining magnification, x100); each picture is representative for nine<br />

mice per group. *, p < 0.05 Mann-Whitney U test; data reproduced in at<br />

least three independent experiments.<br />

500<br />

500<br />

500<br />

250<br />

250<br />

250<br />

0<br />

0<br />

0<br />

CC<br />

C<br />

C<br />

63


change <strong>of</strong> body weight [g]<br />

leptin [pg/ml]<br />

total T3 [ng/dL]<br />

total T4 [µg/dL]<br />

Maren Depke<br />

Results<br />

Liver Gene Expression Pattern in a Mouse Psychological Stress Model<br />

In fact, the increase in circulating glucocorticoids (GCs) was less pronounced than after acute<br />

stress (Depke et al. 2008: supplemental material 1 at http://endo.endojournals.org), but the<br />

continuing hypothalamic-pituitary-adrenal (HPA) axis response might contribute to the reduction<br />

<strong>of</strong> body weight <strong>of</strong> up to 20 % during the time course <strong>of</strong> repeated stress exposure (Fig. R.1.2 A).<br />

This loss <strong>of</strong> body mass was not associated with significantly altered growth hormone (GH)<br />

concentrations in the blood <strong>of</strong> stressed (13.6 ± 7.6 ng/ml) vs. nonstressed mice<br />

(10.8 ± 5.4 ng/ml). However, stress-induced hyperleptinemia was measured (Fig. R.1.2 B). Total<br />

triiodothyronine / T 3 (Fig. R.1.2 C) and total thyroxine / T 4 levels (Fig. R.1.2 D) were reduced in the<br />

plasma <strong>of</strong> stressed animals, whereas thyroglobulin concentrations remained unchanged (data not<br />

shown).<br />

Fig. R.1.2: Repeated psychological stressinduced<br />

loss <strong>of</strong> BW, increase <strong>of</strong> plasma<br />

leptin levels, and hypothyroidism in mice.<br />

A. Loss <strong>of</strong> body mass during the period <strong>of</strong><br />

4.5 days intermittent stress (black box<br />

plots) compared with nonstressed control<br />

mice (white box plots) (n = 12 mice per<br />

group).<br />

B. Plasma leptin levels after nine stress<br />

cycles compared with nonstressed mice<br />

(n = 12 mice per group).<br />

C, D. T 3 (C) and T 4 (D) concentrations in<br />

the plasma <strong>of</strong> repeatedly stressed and<br />

control mice (n = 12 mice per group).<br />

* p < 0.05; ** p < 0.01 Mann-Whitney<br />

U test; data representative for two<br />

independent experiments.<br />

A B C D<br />

*<br />

0<br />

-2.5<br />

-5<br />

*<br />

750<br />

500<br />

250<br />

0<br />

350<br />

*<br />

300<br />

250<br />

200<br />

7.5<br />

5.0<br />

2.5<br />

0<br />

**<br />

Repeated stress-induced changes <strong>of</strong> global hepatic gene expression<br />

In an approach to a more comprehensive <strong>characterization</strong> <strong>of</strong> the metabolic changes that occur<br />

as a result <strong>of</strong> repeated acoustic and restraint stress, the expression signatures <strong>of</strong> liver from<br />

repeatedly stressed BALB/c mice were recorded and compared with those <strong>of</strong> nonstressed<br />

controls. The Affymetrix-based mRNA expression pr<strong>of</strong>iling <strong>of</strong> the liver <strong>of</strong> repeatedly stressed vs.<br />

nonstressed animals revealed induction and repression, respectively, <strong>of</strong> 120 and 50 genes in both<br />

independent stress experiments performed. To discriminate effects <strong>of</strong> repeated stress from those<br />

<strong>of</strong> acute stress, the changes in the hepatic gene expression that occurred as a result <strong>of</strong> a single<br />

stress exposure were additionally analyzed. In this model <strong>of</strong> acute stress, 192 and 123 genes<br />

displayed stress-mediated induction or repression <strong>of</strong> expression.<br />

Comparatively analyzing the effects <strong>of</strong> acute and repeated stress, it became clear that both<br />

types <strong>of</strong> stress target a common set <strong>of</strong> 94 genes. Furthermore, 221 and 76 genes were<br />

predominantly regulated <strong>by</strong> acute and repeated stress, respectively (Fig. R.1.3 A).<br />

To analyze the changes in gene expression within the framework <strong>of</strong> already accumulated<br />

knowledge, the lists <strong>of</strong> genes differentially expressed after acute and repeated stress or both<br />

were subjected to an analysis using the IPA s<strong>of</strong>tware. This s<strong>of</strong>tware allowed for an intuitive<br />

mining <strong>of</strong> the data <strong>of</strong> the 391 differentially expressed genes to gather an impression <strong>of</strong> the<br />

biological rationale <strong>of</strong> the expression changes experimentally observed within the context <strong>of</strong><br />

published data.<br />

When the IPA s<strong>of</strong>tware was used to analyze the molecular and cellular functions targeted <strong>by</strong><br />

stress, an influence on broad categories such as “cell growth and proliferation” and “cell death”<br />

was noted (Fig. R.1.3 B). However, it was also apparent that genes related to metabolic diseases<br />

were most significantly influenced <strong>by</strong> the repeated stress exposure (Fig. R.1.3 C). This finding was<br />

in line with the metabolic disturbances observed before. Supporting this notion <strong>of</strong> a major impact<br />

64


Maren Depke<br />

Results<br />

Liver Gene Expression Pattern in a Mouse Psychological Stress Model<br />

<strong>of</strong> repeated stress on metabolism, highly significant changes were also noted for more specific<br />

categories such as amino acid metabolism and lipid metabolism. Some <strong>of</strong> these influences on<br />

metabolism were already noted during acute stress because changes related to metabolic<br />

disease ranked at number six when genes commonly influenced <strong>by</strong> acute and repeated stress<br />

were analyzed. Genes involved in more specific categories <strong>of</strong> metabolism such as lipid and amino<br />

acid metabolism were only moderately influenced <strong>by</strong> acute stress. Thus, the gene expression<br />

pr<strong>of</strong>iling favors the idea that acute stress sets into motion a gene regulation cascade that is then<br />

manifested during repeated stress exposure finally leading to the observed metabolic<br />

disturbances.<br />

A. Numbers <strong>of</strong> differentially expressed genes after acute and chronic stress or in both models<br />

221<br />

94<br />

76<br />

acute stress specific<br />

differential gene expression<br />

differential gene expression<br />

after both<br />

acute and repeated stress<br />

repeated stress specific<br />

differential gene expression<br />

B. Over-represented functions with highest significance in the lists <strong>of</strong> differentially expressed genes<br />

rank function function function<br />

1 Cell Cycle Cellular Growth and Proliferation Metabolic Disease<br />

2 Cellular Growth and Proliferation Cell Death Hematological Disease<br />

3 Cell Death Connective Tissue Disorders Cell Signaling<br />

4 Gene Expression Inflammatory Disease Immune Response<br />

5 Cellular Development Skeletal and Muscular Disorders Lipid Metabolism<br />

6 Cancer Metabolic Disease Amino Acid Metabolism<br />

C. Selected over-represented metabolic functions in the lists <strong>of</strong> differentially expressed genes<br />

function rank rank rank<br />

Metabolic Disease 26 6 1<br />

Amino Acid Metabolism 24 62 6<br />

Lipid Metabolism 16 16 5<br />

Fig. R.1.3: Impact <strong>of</strong> acute and repeated stress on metabolic genes and genes associated with metabolic disease.<br />

A. Graphical display <strong>of</strong> genes differentially expressed after acute or repeated stress, or both stress types. The numbers given in the<br />

Venn diagram include cDNAs, which are not functionally annotated: 29 <strong>of</strong> 221 genes regulated specifically in acute stress, nine <strong>of</strong> 94<br />

genes regulated in both acute and repeated stress, and two <strong>of</strong> 76 genes regulated specifically for repeated stress.<br />

B, C. The lists <strong>of</strong> genes differentially expressed either only after acute stress and repeated stress or after both acute and repeated<br />

stress were loaded into IPA version 5.5 (Ingenuity Systems) to interpret the affected genes within the context <strong>of</strong> the published<br />

literature. Metabolism seemed to be a major target <strong>of</strong> gene regulation after repeated stress exposure, and, thus, the influence <strong>of</strong><br />

acute and repeated stress on metabolism-associated genes was analyzed. The impact <strong>of</strong> acute or repeated stress alone as well as both<br />

stress types is displayed <strong>by</strong> showing the rankings within the list <strong>of</strong> statistically significantly overrepresented functional groups for<br />

genes related to metabolic disease, amino acid metabolism and lipid metabolism.<br />

B. Over-represented functions with highest significance in the lists <strong>of</strong> differentially expressed genes in acute and repeated stress.<br />

C. Impact <strong>of</strong> acute and repeated stress on metabolic functions.<br />

To elucidate the reasons for stress-induced cachexia, it was decided to concentrate selectively<br />

on changes <strong>of</strong> expression <strong>of</strong> genes whose products are involved in metabolic processes and<br />

regulation <strong>of</strong> metabolic pathways. Several genes that were regulated in the liver <strong>of</strong> repeatedly<br />

stressed animals could be linked to hypercatabolism (Fig. R.1.3). Genes relevant for amino acid<br />

metabolism, especially amino acid transporters and enzymes metabolizing glucogenic amino<br />

65


Maren Depke<br />

Results<br />

Liver Gene Expression Pattern in a Mouse Psychological Stress Model<br />

acids [Slc15a4, Slc25a15, Slc3a1, Asl, and glutamate oxaloacetate transaminase 1 (Got1)], were<br />

mostly induced. Moreover, the liver gene expression pr<strong>of</strong>iling <strong>of</strong> repeatedly stressed mice<br />

indicated increased metabolism <strong>of</strong> lipids [Adh4, Apoa4, Cd74, Chpt1, Cyp17a1, Cyp2b10,<br />

Cyp3a13, Cyp4a10, Cyp8b1, Hsd17b2, Hsd3b2, 4632417N05Rik (Hspc105), Saa2, Slco1a1, Xbp1].<br />

To validate the array data, real-time RT-PCR focusing on chronic stress-induced dysregulation<br />

and its pathophysiological effects was performed. Therefore, genes that were associated with<br />

repeated stress-influenced metabolic processes <strong>of</strong> carbohydrate metabolism (Pck1), fat<br />

metabolism (Srebf1), and amino acid metabolism (Asl, Sds), and with stress-induced apoptosis<br />

(Gadd45b) were chosen. For all selected genes, the regulation found with the Affymetrix-based<br />

expression pr<strong>of</strong>iling was confirmed (Table R.1.1).<br />

Table R.1.1: Real-time PCR validation <strong>of</strong> array data in repeated stress experiments<br />

target gene<br />

control group a<br />

ΔCt (Ct target − Ct reference Actb)<br />

repeated stress group b<br />

p-value<br />

Asl 2.90 ± 0.33 1.50 ± 0.23 < 0.0001<br />

Gadd45b 9.66 ± 0.64 6.83 ± 1.55 < 0.0001<br />

Pck1 2.53 ± 0.52 0.02 ±0.72 < 0.0001<br />

Sds 5.66 ± 0.41 3.70 ± 0.64 < 0.0001<br />

Srebf1 8.01 ±0.18 8.36 ± 0.13 < 0.0001<br />

a Validation <strong>of</strong> expression data <strong>by</strong> real-time PCR was carried out for all individual RNA preparations <strong>of</strong> the two biological experiments<br />

(n = 9 plus n = 8 mice/group) <strong>of</strong> the two experiments focusing on effects <strong>of</strong> repeated stress exposures.<br />

b Differences <strong>of</strong> ΔCt values were analyzed <strong>by</strong> Mann-Whitney test.<br />

Induction <strong>of</strong> gluconeogenesis in repeatedly stressed mice<br />

Stimulated <strong>by</strong> the observed loss in total body mass and the suspected involvement <strong>of</strong><br />

carbohydrate metabolism, the expression pr<strong>of</strong>iles for relevant genes were specifically<br />

investigated, even if this category was not part <strong>of</strong> the first most significant biological functions<br />

according to the IPA categorization. Stress-induced increased expression <strong>of</strong> Foxo1, Igfbp1, Irs1,<br />

and Pck1, as well as reduced mRNA levels <strong>of</strong> Srebf1, can induce hyperglycemia because <strong>of</strong><br />

activation <strong>of</strong> gluconeogenic pathways. In contrast, the gene products <strong>of</strong> Cebpb, Igfbp1, St3gal5,<br />

and Tnfrsf1b are associated with hypoglycemia, and may indicate counterregulatory processes to<br />

decrease blood glucose levels.<br />

In singularly stressed mice, in vivo analysis did not reveal significant changes <strong>of</strong> carbohydrate<br />

regulation pathways, e. g. <strong>of</strong> leptin concentrations in the plasma, blood glucose levels, or liver<br />

histology (Depke et al. 2008: supplemental material 1 at http://endo.endojournals.org).<br />

In contrast, repeated stress induced pathophysiologically relevant alterations <strong>of</strong> protein and<br />

lipid metabolism to provide fuel for gluconeogenesis. Only in chronically stressed mice<br />

disturbances <strong>of</strong> the carbohydrate metabolism became detectable also in vivo. This included a<br />

transient hypoglycemic period immediately after the termination <strong>of</strong> the ninth stress session.<br />

However, after resuming food intake in the home cage, blood glucose levels increased rapidly<br />

and resulted in a prolonged hyperglycemia that still was detectable 2 h later (Fig. R.1.4 A). In the<br />

liver <strong>of</strong> repeatedly stressed mice, an increased usage <strong>of</strong> carbohydrate reservoirs was assessed <strong>by</strong><br />

reduced PAS staining that stains aldehyde groups <strong>of</strong> carbohydrates in tissue and revealed<br />

reduced storage <strong>of</strong> carbohydrates in the liver <strong>of</strong> repeatedly stressed mice compared with healthy<br />

control mice (Fig. R.1.4 B, C). Moreover, after repeated stress, insulin concentrations in the<br />

plasma were slightly increased (272.5 ± 131.4 pg/ml) compared with control mice<br />

(170 ± 149 pg/ml). Resistin, an insulin-resistance inducing adipokine, was significantly increased<br />

in the plasma <strong>of</strong> repeatedly stressed mice when compared with nonstressed animals<br />

(Fig. R.1.4 D). Analysis <strong>of</strong> pH in EDTA plasma samples revealed stress-induced acidosis<br />

(Fig. R.1.4 E).<br />

66


lood glucose levels [nmol/l]<br />

resistin (ng/ml)<br />

pH<br />

Maren Depke<br />

Results<br />

Liver Gene Expression Pattern in a Mouse Psychological Stress Model<br />

A B D E<br />

*<br />

*<br />

15<br />

*<br />

60<br />

*<br />

7.5<br />

*<br />

10<br />

*<br />

C<br />

50<br />

7.4<br />

5<br />

40<br />

7.3<br />

0<br />

0 0.5 2 h<br />

30<br />

7.2<br />

Fig. R.1.4: Disturbances <strong>of</strong> murine carbohydrate metabolism after repeated psychological stress.<br />

A. Kinetics <strong>of</strong> blood glucose levels immediately after termination <strong>of</strong> the ninth stress cycle (black box plots) compared with controls<br />

(white box plot) (n = 9 mice per group).<br />

B, C. Reduced storage <strong>of</strong> carbohydrates in the liver <strong>of</strong> repeatedly stressed mice (B) compared with nonstressed mice (C) (PAS staining<br />

magnification, x200); each picture is representative for nine mice per group.<br />

D, E. Plasma resistin levels (D) and pH <strong>of</strong> EDTA plasma (E) <strong>of</strong> stressed and nonstressed mice (n = 12 mice per group). * p < 0.05 Mann-<br />

Whitney U test; data representative for two independent experiments.<br />

Hypercholesteremia after repeated stress exposure<br />

Global gene expression analysis <strong>of</strong> the liver <strong>of</strong> repeatedly stressed mice revealed stressinduced<br />

changes <strong>of</strong> the gene expression pr<strong>of</strong>ile <strong>of</strong> lipid metabolism (Fig. R.1.3). Therefore, the<br />

lipid turnover <strong>of</strong> these mice was analyzed. After repeated stress exposure, a hepatic steatosis was<br />

observed (Fig. R.1.5 A), whereas no significant numbers <strong>of</strong> lipid vesicles were detected in the liver<br />

<strong>of</strong> control mice (Fig. R.1.5 B). A Sudan III staining, which selectively stains triglycerides but not<br />

cholesterol esters, did not indicate any differences between stressed vs. nonstressed mice (data<br />

not shown). Therefore, the lipids that were accumulated in the liver were presumably not<br />

triglycerides but steroids or their precursor molecules. This is supported <strong>by</strong> the array data that<br />

showed up-regulation <strong>of</strong> genes for steroid metabolism (Cyp17a1, Cyp2b10, Cyp39a1, Cyp4a14,<br />

and Por).<br />

Analysis <strong>of</strong> plasma lipid composition in repeatedly stressed mice showed reduced triglyceride<br />

levels (Fig. R.1.5 C) but increased total cholesterol concentrations (Fig. R.1.5 D). Among<br />

lipoproteins, the HDL fraction was increased (Fig. R.1.5 E), whereas VLDL concentrations were<br />

strongly decreased (Fig. R.1.5 F). LDL-cholesterol levels did not change (Fig. R.1.5 G).<br />

In contrast to the repeated stress model, plasma lipid composition or histological alterations<br />

in the liver were not detected when comparing acutely stressed and control mice (Depke et al.<br />

2008: supplemental material 1 at http://endo.endojournals.org). In addition, the expression<br />

pr<strong>of</strong>iling <strong>of</strong> acutely stressed mice did not reveal major changes in genes involved in lipid<br />

metabolism, probably indicating that hepatocytes <strong>of</strong> stressed mice started an anticipatory gene<br />

expression program during the repeated stress sessions to face the stressful situation whose<br />

physiological impact did not become detectable until stress exposure was repeated.<br />

Loss <strong>of</strong> essential amino acids in repeatedly stressed mice<br />

The gene expression analysis <strong>of</strong> the liver <strong>of</strong> repeatedly stressed animals also showed altered<br />

expression pr<strong>of</strong>iles <strong>of</strong> genes whose products are involved in amino acid metabolism (e. g. Asl,<br />

Got1, Prodh, Slc15a4, Slc25a15, Slc3a1, and Tdo; Fig. R.1.3). Despite the small group size <strong>of</strong><br />

analyzed animals, the amino acid composition <strong>of</strong> fresh plasma samples revealed significantly<br />

67


triglycerides [mmol/l]<br />

cholesterol [mmol/l]<br />

HDL [mmol/l]<br />

VLDL [mmol/l]<br />

LDL [mmol/l]<br />

Maren Depke<br />

Results<br />

Liver Gene Expression Pattern in a Mouse Psychological Stress Model<br />

reduced concentrations <strong>of</strong> several essential amino acids, e. g. arginine, threonine, methionine,<br />

and tryptophan, whereas nonessential amino acids showed fewer alterations in repeatedly<br />

stressed mice (Depke et al. 2008: supplemental material 5 at http://endo.endojournals.org). In<br />

addition, gene expression pr<strong>of</strong>iling <strong>of</strong> the liver <strong>of</strong> repeatedly stressed mice showed an induction<br />

<strong>of</strong> genes for amino acid transporters and amino acid metabolizing enzymes (Sds, Slc15a4,<br />

Slc25a15, Slc3a1, Got1, and Tat), and provided hints for increased activation <strong>of</strong> amino acid<br />

degradation pathways (Aass, Ahcy, Asl, Prodh, and Tdo2). The induction <strong>of</strong> Asl expression along<br />

with a loss <strong>of</strong> arginine and citrulline in the plasma (Depke et al. 2008: supplemental material 5 at<br />

http://endo.endojournals.org) provided hints for altered urea cycle activity. This substantiates<br />

the observations <strong>of</strong> systemic usage <strong>of</strong> the body’s protein stores in repeatedly stressed BALB/c<br />

mice. In contrast, after a single acute stress session, mRNA expression <strong>of</strong> only a few glucogenic<br />

amino acid transporters and metabolizing enzymes was induced in the liver (Sds, Slc38a2, and<br />

Tat), which did not result in altered amino acid levels in the periphery (data not shown).<br />

A<br />

B<br />

C D E F G<br />

4<br />

3.5 **<br />

***<br />

3<br />

3.0<br />

2<br />

2.5<br />

1<br />

2.0<br />

3.0<br />

***<br />

0.3 ***<br />

2.5<br />

0.2<br />

2.0<br />

0.1<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

1.5<br />

1.5<br />

0<br />

0<br />

Fig. R.1.5: Disturbances <strong>of</strong> fat metabolism in repeatedly stressed mice.<br />

A, B. Hepatic steatosis in repeatedly stressed mice (A) compared with nonstressed mice (B) (HE staining magnification, x20); each<br />

picture is representative for nine mice per group.<br />

C–G. Plasma fat composition <strong>of</strong> repeatedly stressed mice (black box plots) and nonstressed controls (white box plots); triglyceride<br />

levels (C), total cholesterol (D), HDL-cholesterol (E), VLDL-cholesterol (F), and LDL-cholesterol (G) were measured immediately after<br />

the ninth stress session (n = 12 mice per group). ** p < 0.01; *** p < 0.001, Mann-Whitney U test.<br />

Stress-induced alterations <strong>of</strong> hepatic gene expression <strong>of</strong> immune response genes after acute<br />

and chronic stress<br />

The analysis <strong>of</strong> liver homogenates <strong>of</strong> acutely and chronically stressed mice revealed an altered<br />

mRNA expression <strong>of</strong> some immune response genes. Here, canonical pathway analysis <strong>of</strong><br />

Ingenuity Pathway Analysis (IPA, www.ingenuity.com) was applied to gain a deeper insight in the<br />

mechanism <strong>of</strong> stress-induced alterations <strong>of</strong> immune functions after acute or repeated<br />

psychological stress. Looking at the effects <strong>of</strong> acute stress, ranking <strong>of</strong> canonical pathways <strong>by</strong> IPA<br />

68


LBP [ng/ml]<br />

CPR [ng/ml]<br />

Maren Depke<br />

Results<br />

Liver Gene Expression Pattern in a Mouse Psychological Stress Model<br />

revealed the following five prime targets: “PXR/RXR activation” (p = 4.04E−06), “Glucocorticoid<br />

Receptor Signaling” (p = 4.26E−05), “LPS/IL-1 Mediated Inhibition <strong>of</strong> RXR function”<br />

(p = 6.44E−05), “IGF-1 Signaling” (p = 7.29E−05), and “Hepatic Fibrosis/Hepatic Stellate Cell<br />

Activation” (p = 2.96E−04). An IPA driven analysis <strong>of</strong> the consequences <strong>of</strong> chronic stress exposure<br />

in turn uncovered a prime influence onto the following five canonical pathways, which partially<br />

overlapped with those influenced <strong>by</strong> acute stress in hepatic tissue: “LPS/IL-1-Mediated Inhibition<br />

<strong>of</strong> RXR Function” (p = 4.38E−06), “PXR/RXR activation” (p = 5.41E−06), “Acute Phase Response<br />

Signaling” (p = 2.48E−05), “Antigen Presentation Pathway” (p = 1.78E−04), and “Glucocorticoid<br />

Receptor Signaling” (p = 4.24E−04).<br />

Already after acute stress, many genes which are associated with immune activation were<br />

induced (e. g. Egfr, Fgf1, Jun, Irf2, Lgals4, Lipg, Map3k5). Several <strong>of</strong> these markers such as Il1r1,<br />

Il6ra, Cxcl1, Lpin1, Tnfrsf1b, or Vcam1 were highly expressed in hepatic tissue also after repeated<br />

stress exposure when compared with non-stressed controls. Interestingly, IPA revealed a high<br />

mRNA expression <strong>of</strong> acute phase response (APR) genes immediately after acute stress exposure<br />

(canonical pathway affected at rank 8 after acute stress: “Acute Phase Response Signaling” with<br />

p = 8.7E−04), which was even further increased after the ninth stress session (Canonical pathway<br />

“Acute phase response”; Table R.1.2). To validate the biological relevance <strong>of</strong> an increased APR,<br />

LPS-binding protein (LBP) and C-reactive protein (CRP) concentrations were determined in<br />

plasma. Acutely stressed mice did not show any significantly increased concentration <strong>of</strong> these<br />

APR proteins after 2 h stress exposure whereas LBP and CRP levels were significantly enhanced in<br />

the plasma <strong>of</strong> chronically stressed mice (Fig. R.1.6).<br />

Fig. R.1.6: Acute phase proteins in the<br />

plasma after acute and chronic stress and<br />

in control mice.<br />

Plasma level <strong>of</strong> (A) LPS-binding protein<br />

(LBP) and (B) C-reactive protein (CRP) <strong>of</strong><br />

acutely stressed mice (grey box plot),<br />

chronically stressed animals (black box<br />

plot), and <strong>of</strong> non-stressed controls.<br />

n = 12 mice/group;<br />

summarized from two independent<br />

experiments with comparable results.<br />

** p < 0.01, *** p < 0.001 <strong>by</strong> one-way<br />

ANOVA and Bonferroni multiple<br />

comparison test.<br />

A<br />

750 *** **<br />

500<br />

250<br />

0<br />

B<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

***<br />

Moreover, gene expression patterns contained several hints for the induction <strong>of</strong> immune<br />

suppressive pathways which included an up-regulation <strong>of</strong> Tsc22d3 (GILZ, glucocorticoid-induced<br />

leucine zipper) or Fkbp5 in both acutely and chronically stressed mice and reduced expression <strong>of</strong><br />

interferon gamma inducible target genes such as Ifi47, Iigp1, Stat1, and Socs2 selectively after<br />

chronic stress compared with acutely stressed mice and non-stressed controls. Importantly,<br />

chronically stressed mice showed a reduced hepatic transcription <strong>of</strong> antigen presentation<br />

pathway molecules such as CD74 antigen (Ia antigen-associated invariant chain) and the<br />

histocompatibility class II antigens H2-Aa and H2-Ea. An overlay <strong>of</strong> the gene expression data to<br />

the pre-defined IPA-canonical pathway “Antigen Presentation Pathway” depicts that chronic<br />

stress-induced repression especially affected genes <strong>of</strong> the MHC class II signaling pathway<br />

(Fig. R.1.7) which may significantly reduce the capacity to mount an antibacterial response.<br />

69


Maren Depke<br />

Results<br />

Liver Gene Expression Pattern in a Mouse Psychological Stress Model<br />

Table R.1.2: Tabular overview <strong>of</strong> stress-regulated hepatic genes, which are included in Ingenuity Pathway Analysis’ (IPA, Ingenuity<br />

Systems, www.ingenuity.com) canonical pathway „Acute Phase Response Signaling“.<br />

information on canonical pathway<br />

„Acute Phase Response Signaling“ in IPA<br />

information on stress regulated hepatic genes<br />

cellular<br />

location and<br />

Function<br />

plasma<br />

membrane<br />

receptor<br />

cytoplasm /<br />

signal<br />

transduction<br />

nucleus /<br />

transcription<br />

regulator<br />

affected<br />

genes <strong>of</strong><br />

acute phase<br />

response<br />

name in<br />

canonical<br />

pathway<br />

(IPA) a<br />

members<br />

in IPA, if<br />

group <strong>of</strong><br />

more than<br />

one gene<br />

gene<br />

name<br />

(NetAffx)<br />

gene title (NetAffx)<br />

probe<br />

set ID<br />

fold change b<br />

in<br />

chronic<br />

stress<br />

in<br />

acute<br />

stress<br />

IL1R1 IL-1R,<br />

Il1r1 interleukin 1 receptor, type I 1448950_at 4.4 4.0<br />

IL-1R/TLR<br />

IL-6R Il6ra interleukin 6 receptor, alpha 1452416_at 3.3 5.5<br />

TNFR NGFR,<br />

TNFRSF1A,<br />

TNFRSF1B,<br />

TNFRSF11B<br />

Tnfrsf1b tumor necrosis factor receptor<br />

superfamily, member 1b<br />

1418099_at 2.4 3.4<br />

ASK1<br />

Map3k5<br />

(LOC<br />

675366)<br />

mitogen activated protein kinase<br />

kinase kinase 5 /// similar to<br />

mitogen activated protein kinase<br />

kinase kinase 5<br />

mitogen activated protein kinase<br />

3<br />

ERK1/2 MAPK1,<br />

MAPK3<br />

Mapk3<br />

GCR* Nr3c1 nuclear receptor subfamily 3,<br />

group C, member 1<br />

IκBα* BCL3,<br />

Nfkbia nuclear factor <strong>of</strong> kappa light<br />

NFKBIA,<br />

chain gene enhancer in B-cells<br />

NFKBIB,<br />

inhibitor, alpha<br />

NFKBIE<br />

SOCS*<br />

SOCS1,<br />

SOCS2,<br />

SOCS3,<br />

SOCS4,<br />

SOCS5,<br />

SOCS6<br />

Socs2<br />

suppressor <strong>of</strong> cytokine signaling<br />

2<br />

1421340_at 2.5<br />

1427060_at -1.6<br />

1460303_at -1.5<br />

1448306_at 1.8 2.0<br />

1449109_at -2.4<br />

c-FOS Fos FBJ osteosarcoma oncogene 1423100_at 2.5<br />

c-JUN Jun Jun oncogene 1448694_at<br />

1417409_at<br />

1.7<br />

2.0<br />

GCR* Nr3c1 nuclear receptor subfamily 3, 1460303_at -1.5<br />

group C, member 1<br />

NF-IL6 Cebpb CCAAT/enhancer binding protein<br />

(C/EBP), beta<br />

1418901_at<br />

1427844_a_at<br />

3.3<br />

4.0<br />

5.1<br />

6.9<br />

CRP Crp C-reactive protein, pentraxinrelated<br />

1421946_at 1.6<br />

IκBα* BCL3,<br />

Nfkbia nuclear factor <strong>of</strong> kappa light 1448306_at 1.8 2.0<br />

NFKBIA,<br />

chain gene enhancer in B-cells<br />

NFKBIB,<br />

inhibitor, alpha<br />

NFKBIE<br />

ORM ORM1,<br />

Orm2 orosomucoid 2 1420438_at 3.3<br />

SAA<br />

SOCS*<br />

ORM2<br />

SAA1,<br />

SAA2,<br />

SAA4<br />

SOCS1,<br />

SOCS2,<br />

SOCS3,<br />

SOCS4,<br />

SOCS5,<br />

SOCS6<br />

Saa1 serum amyloid A 1 1419075_s_at<br />

1450788_at<br />

23.7<br />

27.9<br />

Saa2 serum amyloid A 2 1449326_x_at 30.5<br />

Saa4 serum amyloid A 4 1419318_at<br />

1419319_at<br />

2.6<br />

2.6<br />

Socs2 suppressor <strong>of</strong> cytokine signaling 1449109_at -2.4<br />

2<br />

a Some genes and their products marked <strong>by</strong> * belong to more than one group <strong>of</strong> the first column.<br />

b Fold change values are only given if the gene was considered as regulated in acute or chronic stress experiments.<br />

2.2<br />

2.5<br />

70


Maren Depke<br />

Results<br />

Liver Gene Expression Pattern in a Mouse Psychological Stress Model<br />

Fig. R.1.7: Hepatic repression <strong>of</strong> antigen<br />

presentation related genes after chronic<br />

stress.<br />

Canonical “Antigen Presentation Pathway” <strong>of</strong><br />

Ingenuity Pathway Analysis (adapted from<br />

IPA, Ingenuity Systems, www.ingenuity.com)<br />

displaying genes specifically repressed during<br />

chronic but not during acute stress.<br />

Repression <strong>of</strong> gene expression is indicated in<br />

green. The following genes are affected: H2-<br />

Aa and H2-Ea (MHC-IIα); Cd74 (CLIP); Psmb9<br />

(LMP2). The blue circle indicates gene<br />

repression for H2-Ab1 (MHC-IIβ), which did<br />

not pass the regulation criteria because <strong>of</strong><br />

low expression level.<br />

Increased leukocyte trafficking into the liver <strong>of</strong> chronically stressed mice<br />

Immediately after acute stress, several markers <strong>of</strong> leukocyte migration were differentially<br />

regulated in hepatic tissue compared with non-stressed controls. This set included lymphocyte<br />

chemoattractive molecules such as Cxcl11 and Cxl12, which were repressed, or neutrophils<br />

attractive Cxcl1, which was highly expressed after acute but also after chronic stress exposure. An<br />

induction <strong>of</strong> Vcam1 after acute and chronic stress and <strong>of</strong> Arhgap5 selectively after repeated<br />

stress as well as the repression <strong>of</strong> Bcar3 in the liver after acute stress exposure indicated that the<br />

extravasation <strong>of</strong> leukocytes was facilitated. Additionally, mRNA pr<strong>of</strong>iling uncovered a repression<br />

<strong>of</strong> genes whose products are involved in maintaining the endothelial barrier function such as<br />

claudin 1 (Cldn1), selectively after acute stress, or claudin 2 and 3 after both acute and chronic<br />

stress. To investigate whether these changes in gene expression also caused physiological<br />

changes in leukocyte composition <strong>of</strong> the liver, immun<strong>of</strong>luorescence staining was performed<br />

which revealed that no changes <strong>of</strong> immune cell numbers in the liver were detectable<br />

immediately after a single acute stress session <strong>by</strong> staining leukocytes with anti-CD45 antibodies<br />

(data not shown). However, in the periportal areas <strong>of</strong> chronically stressed animals there were<br />

increased numbers <strong>of</strong> CD45-positive cells, which supports the data <strong>of</strong> the gene expression<br />

pr<strong>of</strong>iling that indicated stress-induced leukocyte migration (data not shown).<br />

Increased oxidative stress in the liver after acute psychological stress exposure is counterregulated<br />

after repeated stress<br />

Inflammatory responses are associated with increased generation <strong>of</strong> reactive oxygen and<br />

nitrogen species. Gene expression pr<strong>of</strong>iling gave hints for increased oxidative stress in the liver <strong>of</strong><br />

acutely and chronically stressed mice. Alas1 or As3mt, which are important for the regulation <strong>of</strong><br />

the oxidative state, were differentially regulated after acute and chronic stress. The redox<br />

71


Maren Depke<br />

Results<br />

Liver Gene Expression Pattern in a Mouse Psychological Stress Model<br />

molecule Nnmt was up-regulated selectively after repeated stress. To verify the biological<br />

relevance <strong>of</strong> an altered redox homeostasis, protein carbonylation was analyzed as a marker <strong>of</strong><br />

oxidative stress. Remarkably, increased protein carbonyl content <strong>of</strong> plasmatic and hepatic<br />

proteins was detectable already immediately after acute stress, indicating a rapid and significant<br />

protein damage <strong>by</strong> increased formation <strong>of</strong> reactive oxygen species (data not shown). After<br />

chronic psychological stress the protein-carbonyl content had returned to the level seen in nonstressed<br />

control mice (data not shown).<br />

Finally, the expression <strong>of</strong> Glrx (glutaredoxin) and Gsta (glutathione S-transferase) was<br />

significantly increased in the liver <strong>of</strong> chronically stressed mice indicating that compensatory antioxidative<br />

mechanisms were induced. This may explain why the carbonyl content <strong>of</strong> proteins in<br />

plasma and liver in repeatedly stressed mice had declined to the normal level observed in nonstressed<br />

control mice (data not shown).<br />

Repeated stress-induced apoptosis in the liver<br />

Several genes that were regulated in the liver <strong>of</strong> repeatedly stressed mice provided hints for<br />

increased cell death. By comparing the mRNA pr<strong>of</strong>ile <strong>of</strong> liver homogenates <strong>of</strong> acutely and<br />

repeatedly stressed mice a highly similar pattern <strong>of</strong> altered gene expression <strong>of</strong> cell cycle and<br />

apoptosis-related genes was observed. In both acutely and chronically stressed animals,<br />

increased mRNA levels compared to non-stressed controls were detected for Tnfrsf1b, Cdkn1a,<br />

Cebpb, Igfbp1, and Gadd45b (Fig. R.1.8 A). Other genes, such as Fos and Jun, which were induced<br />

immediately after acute stress exposure, did not reveal prolonged high mRNA expression<br />

(Fig. R.1.8 A, B), whereas yet another group, including Ccnd1 and Xbp1, were selectively<br />

repressed after the ninth stress session (Fig. R.1.8 B). Using TUNEL techniques, induction <strong>of</strong><br />

hepatocyte apoptosis was not detectable immediately after one single acute stress exposure<br />

(Fig. R.1.8 D), but high numbers <strong>of</strong> dead cells in livers after repeated stress exposure (Fig. R.1.8 E)<br />

compared with control mice (Fig. R.1.8 C) were documented. However, the liver mass after the<br />

ninth stress session was only slightly decreased compared to non-stressed control animals<br />

(Fig. R.1.8 F). This may be caused <strong>by</strong> induction <strong>of</strong> repair processes for which heightened hepatic<br />

expression <strong>of</strong> IL-6 receptor and Ptp4a1 after chronic stress are indications.<br />

Fig. R.1.8. Differential regulation <strong>of</strong> cell death associated genes in liver <strong>of</strong> stressed mice.<br />

Genes resulting from a search for “apoptosis AND liver”, “apoptosis AND hepatocyte”, “cell death AND liver” and “cell death AND<br />

hepatocyte” using Ingenuity Pathway Analysis (IPA, Ingenuity Systems, www.ingenuity.com) were collected in a list. All genes <strong>of</strong> this<br />

list which displayed differential regulation in liver after acute and/or chronic stress exposure were added to a new user-defined<br />

pathway as analysis nodes and <strong>interactions</strong> between these selected genes were drawn using the “Connect” tool <strong>of</strong> IPA. Edge lines that<br />

are continuous represent direct <strong>interactions</strong>, whereas broken lines indicate indirect influences. The pathway was overlaid with<br />

expression data from (A) acute stress and from (B) chronic stress. Red indicates increased and green decreased expression after stress<br />

exposure compared with the control. Intensity <strong>of</strong> the node color represents the degree <strong>of</strong> up (red) and down (green) regulation in the<br />

stressed livers. Different shapes <strong>of</strong> the nodes indicate the functional classes <strong>of</strong> the gene products (e. g. vertical<br />

ellipse − transmembrane receptor; horizontal ellipse − transcription regulator; vertical rectangle − G-protein coupled receptor;<br />

horizontal rectangle − ligand-dependent nuclear receptor; trapezium − transporter; triangle − phosphatase; inverted triangle − kinase;<br />

vertical rhomb − enzyme; horizontal rhomb − peptidase; quadrat − cytokine; double lined circle − group or complex; single lined<br />

circle − other).<br />

The level <strong>of</strong> apoptosis TUNEL assay (x400): compared to the control (C), livers <strong>of</strong> mice exposed to acute stress did no display any<br />

significant increase in the level <strong>of</strong> apoptosis (D). However, high numbers <strong>of</strong> apoptotic hepatocytes were detected in repeatedly<br />

stressed mice (E). Each picture is representative for n = 9 mice/group. Increased apoptosis did not alter liver mass (F), even after the<br />

9th stress cycle (black box plots) compared with non-stressed mice (white box plots) n = 9 mice/group; data representative for least 2<br />

independent experiments.<br />

72


liver weight [g]<br />

Maren Depke<br />

Results<br />

Liver Gene Expression Pattern in a Mouse Psychological Stress Model<br />

A<br />

C<br />

D<br />

B<br />

E<br />

F<br />

1.4<br />

1.2<br />

1<br />

0.8<br />

0.6<br />

73


log cfu/g liver<br />

Maren Depke<br />

Results<br />

Liver Gene Expression Pattern in a Mouse Psychological Stress Model<br />

Reduced antibacterial response in the liver after chronic stress exposure [Cornelia Kiank]<br />

Kiank et al. already showed that chronic psychological stress increased the bacterial spreading<br />

after experimental infection with the extracellular <strong>pathogen</strong> E. coli ATCC 25922 (Kiank et al.<br />

2006). To elucidate whether altered hepatic immune responsiveness <strong>of</strong> chronically stressed<br />

animals also alters the antibacterial response to intracellular microbes, BALB/c mice were<br />

intraperitoneally challenged with 150 cfu Salmonella typhimurium. Salmonella infection could not<br />

be restricted at 48 and 72 h after infection when mice were exposed to nine stress sessions prior<br />

to the bacterial challenge (Fig. R.1.9). This shows that although immune effector cells infiltrated<br />

the liver <strong>of</strong> chronically stressed mice, immune suppression dominated, which resulted in an<br />

insufficient protection against infections.<br />

6<br />

**<br />

***<br />

***<br />

***<br />

5<br />

Fig. R.1.9 Bacterial load <strong>of</strong> livers after infection with Salmonella<br />

typhimurium.<br />

Colony forming units (cfu) in the liver <strong>of</strong> chronically stressed (black<br />

box plots) and non-stressed mice (white box plots) 48 and 72 h<br />

after intraperitoneal infection with 150 cfu <strong>of</strong> S. typhimurium wt<br />

12023 are displayed. (n = 9 mice/group, ** p < 0.01, *** p < 0.001<br />

<strong>by</strong> Mann-Whitney U-test, data representative for two independent<br />

experiments).<br />

4<br />

3<br />

48 h 72 h<br />

74


infection rate<br />

cfu / 10 mg tissue<br />

cfu / 10 mg tissue<br />

Maren Depke<br />

Results<br />

KIDNEY GENE EXPRESSION PATTERN IN AN<br />

IN VIVO INFECTION MODEL<br />

Infection rate in kidney samples<br />

The two Staphylococcus aureus strains RN1HG and RN1HG ΔsigB were used to infect mice<br />

with an almost equal infection dose for both strains. It was known from other genetic<br />

backgrounds that virulence and bacterial load are <strong>of</strong>ten similar for sigB deletion mutants and<br />

their parental strain. Nevertheless, individual mice might still display differences. Therefore, the<br />

infection rate <strong>of</strong> each sample was tested on molecular basis to identify potentially existing outlier<br />

samples <strong>of</strong> divergent bacterial load.<br />

Infection rates <strong>of</strong> individual mice kidney samples were comparable in range, mean, and<br />

median in the biological replicates (BR) samples as well used as in for samples 10 arrays_Exp <strong>of</strong> infection Feb 09with the two different<br />

strains (Fig. R.2.1). The infection rate <strong>of</strong> samples and infected exp. Apr 2009 with S. aureus RN1HG ranged from<br />

4.3E+05 cfu/10 mg tissue to 2.2E+06 cfu/10 mg tissue (mean: 1.1E+06; median: 9.2E+05) and <strong>of</strong><br />

those infected with RN1HG ΔsigB from 4.7E+05 cfu/10 mg tissue to 2.3E+06 cfu/10 mg tissue<br />

Infection rate cfu / 10 mg tissue<br />

(mean: 1.0E+06; median: 8.6E+05) in both biological [experiments replicates. february and april 2009]<br />

3.0E+06 3.010 6<br />

Fig. R.2.1:<br />

Infection rate in kidney samples <strong>of</strong><br />

mice infected with S. aureus RN1HG<br />

and RN1HG ΔsigB.<br />

The infection rate <strong>of</strong> each sample was<br />

determined in a qPCR approach in<br />

comparison to data from a mixture <strong>of</strong><br />

non-infected kidney tissue and in vitro<br />

cultivated staphylococcal cells. The<br />

line indicates the mean <strong>of</strong> infection<br />

rate for each biological replicate.<br />

BR – biological replicate.<br />

2.0E+06 2.010 6<br />

1.0E+06 1.010 6<br />

infection with<br />

RN1HG<br />

first biological<br />

replicate BR1<br />

wt (2 kidneys)_Feb09<br />

NMRI infection infected with with infection RN1HGwith<br />

RN1HG ΔsigB RN1HG<br />

first biological second biological<br />

replicate BR1 replicate BR2<br />

delta sigB (2 kidneys)_Feb09<br />

wt (2 kidneys)_Apr09<br />

infection with<br />

RN1HG ΔsigB<br />

second biological<br />

replicate BR2<br />

delta sigB (2 kidneys)_Feb09<br />

Reproducibility <strong>of</strong> replicates and clustering <strong>of</strong> Exp. treatment Feb. 09; group members Exp. Apr. 09;<br />

used for array analysis (4/7; 5/8) used for array analysis (5/8; 5/8)<br />

Principal Component Analysis (PCA) was applied for a first general impression <strong>of</strong> the array<br />

data set. This method calculates the direction <strong>of</strong> strongest variation from the multidimensional<br />

array data set, and reduces it to a new value <strong>of</strong> the parameter called Principle Component (PC).<br />

The remaining variation in the data set is subsequently addressed in the same way until all or a<br />

pre-defined fraction <strong>of</strong> variation is collapsed into new values. This procedure results in a set <strong>of</strong><br />

PCs, each <strong>of</strong> which accounts for a fraction <strong>of</strong> the total variance in the data set. Usually, the first 2<br />

or 3 PCs are displayed in a 2- or 3- dimensional plot, respectively. In such a plot, the distance <strong>of</strong><br />

the points that represent the individual data sets correlates to the difference between them.<br />

75


Maren Depke<br />

Results<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

In this study, the PCA plot was derived from log-transformed intensity data <strong>of</strong> 24 arrays,<br />

analyzing 2 biological replicates <strong>of</strong> mice infected with S. aureus RN1HG and RN1HG ΔsigB and one<br />

additional group <strong>of</strong> sham infection (Fig. R.2.2). The analysis was performed on sequence level<br />

(probe sets). Control sequences and sequences that are absent on all 24 arrays (p > 0.01 on<br />

intensity pr<strong>of</strong>ile level) were not included.<br />

The biological reproducibility is visualized in the PCA plot <strong>by</strong> the arrangement <strong>of</strong><br />

corresponding data points close to each other. The PCA clearly distinguished two groups:<br />

infection with S. aureus and sham infection / NaCl control (Fig. R.2.2 A). Furthermore, the PCA<br />

depicted that the data sets <strong>of</strong> infection with S. aureus RN1HG and infection with RN1HG ΔsigB<br />

were highly similar (Fig. R.2.2 A−C).<br />

A view from front B view from right side C view from above<br />

Fig. R.2.2: Visualization <strong>of</strong> transcriptome data using Principal Component Analysis (PCA).<br />

Log-transformed data were used to derive a PCA plot containing 24 arrays (analyzing 2 biological replicates <strong>of</strong> 2 groups with different<br />

infecting strains and one additional group <strong>of</strong> sham infection). Resulting principal components (PC) were set to cover 95 % <strong>of</strong> total<br />

variation, while the total number <strong>of</strong> principal components was not limited. The analysis was restricted to non-control probe sets that<br />

were not absent on all 24 arrays, when absence <strong>of</strong> expression is defined via a p-value > 0.01 on intensity pr<strong>of</strong>ile level in Rosetta<br />

Resolver.<br />

All 3 images (A-C) belong to the same PCA shown in the normal view from the front in the first image (A). In the second image the<br />

PCA-graph is turned <strong>by</strong> 90° to the left around the axis <strong>of</strong> principle component 2 resulting in a view from the right side into the plot (B).<br />

Finally, in the third image the view from above into the graph is achieved <strong>by</strong> turning the plot <strong>by</strong> 90° around the axis <strong>of</strong> principle<br />

component 1 (C).<br />

Arrays <strong>of</strong> the first biological replicate (BR1) are represented <strong>by</strong> rhombi () and arrays <strong>of</strong> the second biological replicate (BR2) <strong>by</strong><br />

dots (•). Coloring distinguishes the three treatment groups <strong>of</strong> infection with S. aureus RN1HG (red), infection with S. aureus<br />

RN1HG ΔsigB (blue) and sham infection / NaCl control (green).<br />

Comparison <strong>of</strong> treatment groups reveals the same <strong>host</strong> reaction to infection with S. aureus<br />

RN1HG and its sigB-mutant<br />

Already the PCA had shown a high similarity between the reaction to the two different<br />

infecting strains S. aureus RN1HG and RN1HG ΔsigB. Based on the PCA results the strategy for<br />

statistical testing consisted <strong>of</strong> four main types <strong>of</strong> comparison (Fig. R.2.3):<br />

comparison <strong>of</strong> the two groups with different infecting strains (infection with RN1HG ΔsigB<br />

vs. infection with RN1HG) after combining both biological replicates<br />

comparison <strong>of</strong> the two groups with different infecting strains for each biological replicate<br />

separately<br />

comparison <strong>of</strong> the same experimental groups <strong>of</strong> different biological replicates (infection with<br />

RN1HG: BR1 vs. BR2; infection with RN1HG ΔsigB: BR1 vs. BR2)<br />

comparison <strong>of</strong> infection with S. aureus and sham infection in the second biological replicate<br />

(infection with RN1HG vs. sham infection; infection with RN1HG ΔsigB vs. sham infection)<br />

76


Maren Depke<br />

Results<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

infection with<br />

RN1HG<br />

BR1<br />

(4 arrays)<br />

3<br />

infection with<br />

RN1HG<br />

BR2<br />

(5 arrays)<br />

4<br />

2<br />

1<br />

2<br />

sham infection<br />

(NaCl)<br />

BR2<br />

(5 arrays)<br />

infection with<br />

RN1HG ΔsigB<br />

BR1<br />

(5 arrays)<br />

3<br />

infection with<br />

RN1HG ΔsigB<br />

BR2<br />

(5 arrays)<br />

4<br />

Fig. R.2.3: Overview on the comparisons between groups in this study that were addressed with statistical testing and visualized with<br />

scatter plots.<br />

The two groups with different infecting strains (infection with RN1HG ΔsigB vs. infection with RN1HG) were compared after<br />

combining both biological replicates () and for each biological replicate separately (). Same experimental groups <strong>of</strong> different<br />

biological replicates (infection with RN1HG: BR1 vs. BR2; infection with RN1HG ΔsigB: BR1 vs. BR2) were also checked against each<br />

other (). The last comparison focused on infection and sham infection in the second biological replicate (infection with RN1HG vs.<br />

sham infection; infection with RN1HG ΔsigB vs. sham infection; ).<br />

BR – biological replicate.<br />

The comparisons were visualized using scatter plots (Fig. R.2.4). Comparable to the PCA<br />

results, a striking concordance between expression values <strong>of</strong> kidney after infection with RN1HG<br />

and infection with RN1HG ΔsigB was observed, especially when both biological replicates were<br />

combined (Fig. R.2.4, panel ), but also when the biological replicates were considered<br />

separately (Fig. R.2.4, panel ). In the comparison <strong>of</strong> the same experimental groups in both<br />

replicates high similarity was observed (Fig. R.2.4, panel ), although the inter-replicate variation<br />

<strong>of</strong> the same treatment was higher than the intra-replicate variation <strong>of</strong> the different treatments<br />

for S. aureus infected samples. The inter-replicate variation might be due to the difference <strong>of</strong> one<br />

day in the sampling time point (d 4 or d 5). Strong effects <strong>of</strong> S. aureus infection independent <strong>of</strong><br />

the infecting strain emerged in the comparison to the sham infected / NaCl control sample group<br />

(Fig. R.2.4, panel ).<br />

The different experimental groups were compared with statistical testing to obtain lists <strong>of</strong><br />

differentially expressed genes. Sequences not expressed and control sequences were not<br />

included in statistical testing.<br />

The first statistical test compared kidney samples <strong>of</strong> mice infected with S. aureus RN1HG ΔsigB<br />

vs. infection with RN1HG in an approach that combined both biological replicates (for<br />

comparison see Fig. R.2.3; and Fig. R.2.4, panel ). Only one sequence corresponding to one<br />

gene was significantly different in intensity between both groups (Table R.2.1). For all arrays<br />

except one array from a specific animal the signal intensities <strong>of</strong> this gene were low and their p-<br />

value for expression was high. i. e. the expression was absent (Fig. R.2.5 A). The array data from<br />

the single outlying animal caused statistical significance, but the result is not biologically relevant<br />

because it is obviously due to an unknown, animal-specific factor and not to treatment.<br />

77


BR2: inf. RN1HG<br />

BR2: inf. RN1HG<br />

BR1: inf. RN1HG ΔsigB<br />

BR1+2: inf. RN1HG ΔsigB<br />

BR2: inf. RN1HG ΔsigB<br />

BR2: inf. RN1HG ΔsigB<br />

BR2: inf. RN1HG ΔsigB<br />

Maren Depke<br />

Results<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

<br />

BR1+2: inf. RN1HG<br />

<br />

BR1: inf. RN1HG<br />

BR2: inf. RN1HG<br />

<br />

BR1: inf. RN1HG<br />

BR1: inf. RN1HG ΔsigB<br />

<br />

BR2: sham inf. (NaCl)<br />

BR2: sham inf. (NaCl)<br />

Fig. R.2.4: Scatter plots comparing mean signal intensities <strong>of</strong> treatment groups.<br />

The signals <strong>of</strong> the three groups “infection with RN1HG”, “infection with RN1HG ΔsigB”, and “sham infection / NaCl control” are plotted<br />

separately for biological replicates (BR1, BR2) or combined for both biological replicates (BR1+2). Control sequences and sequences<br />

that were absent on all <strong>of</strong> the arrays used for each scatter plot are not shown. (Absence <strong>of</strong> expression is defined via a<br />

p-value > 0.01 on intensity pr<strong>of</strong>ile level in Rosetta Resolver). Numbers in the first column refer to the comparison scheme in Fig. R.2.3.<br />

78


Table R.2.1: Genes displaying statistically different expression values in the corresponding comparisons.<br />

a<br />

The GeneChip Mouse Gene 1.0 ST array contains 35557 sequences (probe sets) <strong>of</strong> which 6613 are controls (e. g. negative, positive, PolyA, and hybridization controls). The control sequences and sequences that<br />

were not expressed on all arrays in the selected group comparison were not included in statistical testing.<br />

Maren Depke<br />

Results<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

b Rosetta Resolver s<strong>of</strong>tware maps the sequences (probe sets) to 20074 EntrezGene records (genes). This mapping is used to calculate the numbers <strong>of</strong> genes that are significant in statistical testing and when indicated<br />

group comparison<br />

number<br />

<strong>of</strong><br />

arrays<br />

number <strong>of</strong><br />

sequences<br />

absent on<br />

all arrays<br />

significant with p*


Maren Depke<br />

Results<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

In the search for specific differences in infection with S. aureus RN1HG ΔsigB vs. infection with<br />

RN1HG, the same comparison was performed for the biological replicates separately (for<br />

comparison see Fig. R.2.3; and Fig. R.2.4, panel ). This analysis resulted in 7 or 6<br />

sequences/genes which were differentially expressed in the first biological replicate BR1 or the<br />

second biological replicate BR2, respectively (Table R.2.1). Both lists did not display any overlap<br />

indicating that the few regulated genes were not reproducible in the biological replicates,<br />

although statistically significant (Fig. R.2.5 B).<br />

Additionally, the statistical significance in the biological replicate BR1 was again caused <strong>by</strong> one<br />

single sample, i. e. <strong>by</strong> one specific animal, which provided an outlier value to the data set<br />

(Fig. R.2.5 C). The same sample accounted for the outlier in the values <strong>of</strong> the single gene<br />

significantly regulated when combining the biological replicates (Fig. R.2.5 A).<br />

The statistical significance <strong>of</strong> the 6 differentially expressed genes in the second biological<br />

replicate was not caused <strong>by</strong> outliers (Fig. R.2.5 D), but the fold change in this comparison was<br />

only moderate ranging from to -1.34 to -2.18 (mean: -1.72; median: -1.67).<br />

In conclusion, the study could not provide any hints for statistically significant differences in<br />

the expression pattern in murine kidney upon infection with S. aureus RN1HG or its isogenic sigB<br />

mutant.<br />

When asking for the reproducibility <strong>of</strong> the same treatment in both biological replicates in the<br />

light <strong>of</strong> a possible influence on the detection <strong>of</strong> differential gene expression between the groups<br />

infected with the two different S. aureus strains (for comparison see Fig. R.2.3; and Fig. R.2.4,<br />

panel ), 33 sequences (i. e. 26 genes) were significantly different between the replicates <strong>of</strong><br />

infection with RN1HG, and 62 sequences (i. e. 31 genes) significantly differed between the<br />

replicates <strong>of</strong> infection with RN1HG ΔsigB (Table R.2.1).<br />

Of these genes, only a small fraction <strong>of</strong> approximately 20 % possessed an absolute fold change<br />

greater than 2. For these few genes a high expression variation within one biological replicate<br />

was observed or the statistical difference even was again due to one outlier array. Additionally,<br />

most genes which differed between the replicates did not display any overlap with the genes<br />

which were significantly different between the groups infected with the two different S. aureus<br />

strains. Only some genes were statistically differentially expressed between the equally treated<br />

replicates as well as between infection with RN1HG and RN1HG ΔsigB in the first biological<br />

replicate. But as the statistical significance between samples infected with the two strains was<br />

only caused <strong>by</strong> one single outlier value (Fig. R.2.5 C), it was clear that the small difference<br />

between biological replicates did not prevent the identification <strong>of</strong> differential gene expression in<br />

the comparison <strong>of</strong> <strong>host</strong> reaction to the two infecting S. aureus strains.<br />

In summary, the comparison <strong>of</strong> same treatment groups in different biological replicates<br />

revealed minor, negligible differences that did not influence the detection <strong>of</strong> differential<br />

regulation when comparing the groups <strong>of</strong> infection with different S. aureus strains. The identical<br />

reaction to infection with S. aureus RN1HG and S. aureus RN1HG ΔsigB was reliably measured in<br />

the array study and the detection <strong>of</strong> differences was not prevented <strong>by</strong> biological variation<br />

between the two independent infection experiments.<br />

80


Maren Depke<br />

BR1 sign DsigB vs wt<br />

BR1 sign DsigB vs wt<br />

A<br />

DsigB<br />

Igk-V28<br />

wt<br />

infection with RN1HG ΔsigB vs. RN1HG;<br />

BR1 and und BR2 BR2 combined DsigB vs wt<br />

Igk-V28_DsigB<br />

Results<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

B<br />

BR1 sign DsigB vs wt<br />

BR1 sign DsigB vs wt<br />

7 0 6<br />

C<br />

Igk-V28_DsigB<br />

1 10 100 1000 10000<br />

Igk-V28_wt<br />

signal intensity<br />

infection with S. aureus RN1HG ΔsigB<br />

infection with S. aureus RN1HG wt<br />

Igk-V28_wt<br />

Cyp11b1_DsigB<br />

D<br />

sequences/genes differentially<br />

expressed between<br />

infection with RN1HG ΔsigB<br />

and infection with RN1HG<br />

in BR1<br />

sequences/genes differentially<br />

expressed between<br />

infection with RN1HG ΔsigB<br />

and infection with RN1HG<br />

in BR2<br />

Cyp11b1_DsigB<br />

infection BR1 with sign RN1HG DsigB ΔsigB vs wt vs. RN1HG; BR1<br />

Cyp11b1_wt<br />

infection with S. aureus RN1HG ΔsigB<br />

infection with S. aureus RN1HG wt<br />

Igk-V28_DsigB<br />

Igk-V28<br />

Cyp11b1_wt<br />

Igk-V28_wt<br />

4921521F21Rik_DsigB<br />

infectionBR2 withsign RN1HG DsigB ΔsigBvs vs. wt RN1HG; BR2<br />

Cyp11b1_DsigB<br />

Cyp11b1<br />

4921521F21Rik_DsigB<br />

Cyp11b1_wt<br />

4921521F21Rik_wt<br />

Gpr84_DsigB<br />

Gpr84<br />

Gpr84_wt<br />

4921521F21Rik_DsigB<br />

4921521F21Rik<br />

4921521F21Rik_wt<br />

4921521F21Rik_wt<br />

Star_DsigB<br />

Cd274_DsigB<br />

Cd274<br />

Cd274_wt<br />

Star_DsigB<br />

Star<br />

Star_wt<br />

Star_DsigB<br />

Star_wt<br />

Cxcl9_DsigB<br />

Cxcl9<br />

Cxcl9_wt<br />

Hsd3b1_DsigB<br />

Cybb_DsigB<br />

Hsd3b1<br />

Hsd3b1_wt<br />

Star_wt<br />

Hsd3b1_DsigB<br />

Cybb<br />

Cybb_wt<br />

Cyp21a1_DsigB<br />

Cyp21a1<br />

Cyp21a1_wt<br />

Hsd3b1_DsigB<br />

Hsd3b1_wt<br />

C3ar1_DsigB<br />

C3ar1<br />

C3ar1_wt<br />

Cyp11a1_DsigB<br />

Cyp11a1<br />

Cyp11a1_wt<br />

Hsd3b1_wt<br />

Cyp21a1_DsigB<br />

Plekho2_DsigB<br />

Plekho2<br />

Plekho2_wt<br />

1 10 100 1000 10000<br />

signal intensity<br />

Cyp21a1_DsigB<br />

Fig. R.2.5: Signal intensities Cyp21a1_wt and list comparison for differentially expressed genes.<br />

A. Comparison <strong>of</strong> murine kidney expression pr<strong>of</strong>iles after infection with RN1HG ΔsigB and infection with RN1HG with statistical testing<br />

results in one differentially expressed gene when combining both available biological replicates. However, the signal intensity is<br />

extremely low (not expressed with p > 0.01 on intensity pr<strong>of</strong>ile level in Rosetta Resolver s<strong>of</strong>tware) in all arrays except in one array<br />

(i. e. a kidney sample) Cyp21a1_wt from one mouse infected with RN1HG (BR1) which is visible as outlier and which causes the statistical<br />

significance. Cyp11a1_DsigB<br />

B. Comparison <strong>of</strong> genes differentially regulated between infection with RN1HG and its isogenic sigB mutant in BR1 and BR2. Both<br />

10 biological replicates result 100in completely different 1000 lists <strong>of</strong> differentially expressed 10000 genes.<br />

C, D. Signal intensity values <strong>of</strong> differentially expressed genes in the comparison <strong>of</strong> infection with RN1HG ΔsigB and infection with<br />

RN1HG in BR1 Cyp11a1_DsigB<br />

(C) and in BR2 (D). In BR1, statistical significance is caused <strong>by</strong> a single array that includes an outlier value.<br />

Cyp11a1_wt<br />

BR – biological replicate.<br />

10 100 1000 10000<br />

81<br />

1 10 100 1000 10000<br />

signal intensity<br />

Cyp11a1_wt1 10 100 1000 10000<br />

1 10 100 1000 10000


log 2 ratio (inf. RN1HG ΔsigB / sham inf.)<br />

Maren Depke<br />

Results<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

Comparative analysis <strong>of</strong> S. aureus infected samples with sham infection identifies strong<br />

infection/inflammation reactions <strong>of</strong> the kidney tissue<br />

As high reproducibility <strong>of</strong> kidney expression pr<strong>of</strong>iles in biological replicates <strong>of</strong> infection with<br />

S. aureus had been shown, the comparison between S. aureus infected samples and sham<br />

infection / NaCl control was conducted with only one representative biological replicate (for<br />

comparison see Fig. R.2.3; and Fig. R.2.4, panel ).<br />

When comparing sham infection with infection with S. aureus RN1HG, 5264 sequences were<br />

differentially expressed which correspond to 4613 genes (EntrezGene records in Rosetta Resolver<br />

s<strong>of</strong>tware). Of these genes, 1083 possessed an absolute fold change <strong>of</strong> at least 2. A similar<br />

difference was observed in the comparison <strong>of</strong> sham infection with infection with S. aureus<br />

RN1HG ΔsigB: 4826 sequences (4248 genes) differed significantly <strong>of</strong> which 970 genes exhibited<br />

an absolute fold difference <strong>of</strong> at least 2 (Table R.2.1).<br />

When calculating ratios <strong>of</strong> highly similar sample data sets to a common baseline, a result <strong>of</strong><br />

conforming values is expected. The log-ratio plot <strong>of</strong> the two ratios “infection with<br />

RN1HG” / “sham infection” and “infection with RN1HG ΔsigB” / “sham infection” in the second<br />

biological replicate BR2 on EntrezGene level visualizes this similarity <strong>of</strong> the ratio data: Data points<br />

are mainly arranged near the diagonal <strong>of</strong> the plot (Fig. R.2.6). Some data points display more<br />

scattering, but the difference in the ratios was not significant when tested statistically and was<br />

mainly due to higher variation in signal intensities <strong>of</strong> one group, as described above.<br />

6<br />

Fig. R.2.6:<br />

Ratio plot comparing the log 2ratio <strong>of</strong> “infection with<br />

RN1HG” / “sham infection” with the log 2ratio <strong>of</strong><br />

“infection with RN1HG ΔsigB” / “sham infection” in the<br />

second biological replicate BR2 on EntrezGene level.<br />

Genes differentially regulated (p* < 0.01 in errorweighted<br />

one-way ANOVA with Benjamini-Hochberg<br />

False Discovery Rate multiple testing correction in the<br />

Rosetta Resolver s<strong>of</strong>tware) in both comparisons are<br />

colored in light gray, genes specifically significant in the<br />

comparison “infection with RN1HG” / “sham infection”<br />

are depicted in black while genes specifically significant<br />

in the comparison “infection with RN1HG<br />

ΔsigB” / “sham infection” are shown in dark gray. In<br />

total, 5025 genes are visible.<br />

BR – biological replicate.<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-6 -4 -2 0 2 4 6<br />

log 2 ratio (inf. RN1HG / sham inf.)<br />

The genes differentially regulated in S. aureus infected kidney tissue compared to sham<br />

infection were further analyzed using the Ingenuity Pathway Analysis tool. Only genes displaying<br />

an at least tw<strong>of</strong>old difference in expression between sham infected and S. aureus infected<br />

animals (either RN1HG or RN1HG ΔsigB from the second biological replicate BR2) were<br />

considered in this analysis. Using this approach, the list <strong>of</strong> genes eligible for analysis was<br />

restricted to 1156 genes <strong>of</strong> which 4 could not be mapped to internal data base entries (Ingenuity<br />

Pathway Knowledge Base, IPKB) <strong>by</strong> IPA. Since expression values <strong>of</strong> experiments involving<br />

infection with S. aureus RN1HG or its isogenic sigB mutant did not differ significantly, the average<br />

<strong>of</strong> both expression values was compared to sham infected animals, which allows intuitive<br />

visualization <strong>of</strong> the data.<br />

The global functional analysis using IPA <strong>of</strong>fers an overview on the biological functions that are<br />

associated with the analyzed data set. As expected from the experimental setting, influence on<br />

the functional category “Inflammatory Response” was most pronounced with p-values <strong>of</strong><br />

2.36E−75 to 1.24E−10 for the sub-categories and with 343 associated genes differentially<br />

expressed between infection and sham infection. Additionally, other infection, immune response<br />

82


Maren Depke<br />

Results<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

and inflammation related categories (e. g. Inflammatory / Immunological / Infectious Disease,<br />

Antigen Presentation) and categories showing the impact <strong>of</strong> infection on the tissue (e. g. Cellular<br />

Compromise, Cell Death) were significantly affected.<br />

More detailed analysis <strong>of</strong> the tissue reaction is <strong>of</strong>fered <strong>by</strong> the so-called Canonical Pathway<br />

analysis in IPA. This tool contains predefined pathways and the associated genes, displayed in<br />

order <strong>of</strong> cellular location and function, which can be overlaid with gene expression data, e. g. fold<br />

change values. A p-value for enrichment <strong>of</strong> pathway members in the data set <strong>of</strong> interest<br />

(compared to a reference set, e. g. the whole array) is calculated. Many pathways scored with a<br />

highly significant p-value because <strong>of</strong> the large input data set. Additionally, a ratio <strong>of</strong> the<br />

pathway’s genes included in the data set and the total number <strong>of</strong> genes in the pathway is given.<br />

When data <strong>of</strong> such a Canonical Pathway analysis were ordered <strong>by</strong> p-value, “Acute Phase<br />

Response Signaling” with p = 5.00E−21 and a ratio <strong>of</strong> 53/178 <strong>of</strong> genes from the data set included<br />

in the pathway was the most significantly influenced reaction. This pathway includes signaling<br />

chains starting with TNF-α, IL-1 and IL-6, their signal transduction molecules and transcription<br />

factors and finally the genes whose transcription is induced or repressed. While the acute phase<br />

response is located in hepatic tissue the pathway has a high overlap with a localized<br />

infection/inflammation reaction and therefore is covered <strong>by</strong> the kidney infection vs. sham<br />

infection data set to this high extent.<br />

Directly infection related genes are included in the canonical pathways “Role <strong>of</strong> Pattern<br />

Recognition Receptors in Recognition <strong>of</strong> Bacteria and Viruses” (p = 1.16E−13; ratio = 27/80,<br />

Table R.2.2) and “Toll-like Receptor Signaling” (p = 7.72E−08; ratio = 16/54, Fig. R.2.7) which<br />

contribute as part <strong>of</strong> the innate immune system to the first line <strong>of</strong> defense against <strong>pathogen</strong>s. In<br />

infected kidney tissue, an induction <strong>of</strong> different Toll-like receptors (Tlr1, Tlr2, Tlr4, Tlr6, Tlr7, Tlr8),<br />

the bacterial or dsRNA receptor NALP3 (Nlrp3), and the beta-glucan receptor Dectin-1 (Clec7a)<br />

was observed, and a tendency <strong>of</strong> induction in Tlr9 and the dsRNA receptors RIG-1 (Ddx58) and<br />

Ifih1. Functionally associated receptors CD14 and LBP were also induced and, following infection,<br />

even members <strong>of</strong> the signal transduction cascade showed an increased expression (Syk, Myd88,<br />

Irak3, Map3k1, Irf7, Pik3cd, Pik3cg, Pik3r5, Casp1, Fos, Jun, Nfkb2, and a tendency <strong>of</strong> increase in<br />

Irak4, Mapk13, Map4k4 and Nfkb1). Contrarily, the MAP-kinase-pathway component Map2k6<br />

was repressed and the signal transducer Ecsit and transcription factor PPARα (Ppara) were<br />

repressed <strong>by</strong> trend. The signaling ends in the transcription <strong>of</strong> pro-inflammatory cytokines, e. g.<br />

TNFα, IL-6, and RANTES (Ccl5) whose induction was recorded in infected tissue.<br />

Fig. R.2.7:<br />

Toll-like Receptor Signaling (modified<br />

from IPA, www.ingenuity.com).<br />

Colors indicate direction and magnitude<br />

<strong>of</strong> differential regulation: red – induction;<br />

green – repression. More intense shade <strong>of</strong><br />

color is used for higher absolute fold<br />

change values. A trend <strong>of</strong> induction is<br />

shown in yellow, a trend <strong>of</strong> repression in<br />

light blue.<br />

83


Maren Depke<br />

Results<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

Table R.2.2: Tabular overview <strong>of</strong> infection-regulated genes which are included in Ingenuity Pathway Analysis (IPA, Ingenuity Systems,<br />

www.ingenuity.com) canonical pathway “Role <strong>of</strong> Pattern Recognition Receptors in recognition <strong>of</strong> Bacteria and Viruses” (modified<br />

from IPA, www.ingenuity.com).<br />

IPA<br />

Rosetta Resolver annotation<br />

fold change<br />

infection vs.<br />

sham infection<br />

localization<br />

and functional<br />

grouping a<br />

gene<br />

name<br />

gene<br />

name<br />

description<br />

EntrezGene<br />

ID<br />

RN1HG<br />

RN1HG<br />

ΔsigB<br />

extracellular PRR C1q C1qa complement component 1, q<br />

12259 5.5 4.5<br />

subcomponent, alpha polypeptide<br />

C1qb complement component 1, q<br />

12260 7.4 6.0<br />

subcomponent, beta polypeptide<br />

C1qc complement component 1, q<br />

12262 6.9 5.4<br />

subcomponent, C chain<br />

C3 C3 complement component 3 12266 5.1 5.1<br />

membranebound<br />

C3aR C3ar complement component 3a receptor 1 12267 10.3 6.7<br />

PRR C5aR C5ar complement component 5a receptor 1 12273 7.6 6.1<br />

TLR1 Tlr1 toll-like receptor 1 21897 2.9 2.3<br />

TLR2 Tlr2 toll-like receptor 2 24088 4.4 4.0<br />

TLR4 Tlr4 toll-like receptor 4 21898 2.4 2.3<br />

TLR6 Tlr6 toll-like receptor 6 21899 2.4 2.0<br />

TLR7 Tlr7 toll-like receptor 7 170743 3.0 2.8<br />

TLR8 Tlr8 toll-like receptor 8 170744 4.7 4.3<br />

TLR9 Tlr9 toll-like receptor 9 81897 1.9 1.9<br />

DECTIN-1 Clec7a C-type lectin domain family 7, member a 56644 3.7 2.8<br />

cytoplasmic PRR NALP3 Nlrp3 NLR family, pyrin domain containing 3 216799 3.3 2.7<br />

RIG-1 Ddx58 DEAD (Asp-Glu-Ala-Asp) box polypeptide 58 230073 1.7 1.6<br />

MDA-5 Ifih1 interferon induced with helicase C domain 1 71586 1.5 1.3<br />

signal<br />

PI3K Pik3cd phosphatidylinositol 3-kinase catalytic delta 18707 2.5 2.0<br />

transduction,<br />

effector<br />

Pik3cg<br />

polypeptide<br />

phosphoinositide-3-kinase, catalytic, gamma 30955 3.4 3.1<br />

molecules<br />

(cytoplasm)<br />

Pik3r5<br />

polypeptide<br />

phosphoinositide-3-kinase, regulatory<br />

320207 2.4 2.0<br />

subunit 5, p101<br />

MYD88 Myd88 myeloid differentiation primary response<br />

17874 2.8 2.5<br />

gene 88<br />

SYK Syk spleen tyrosine kinase 20963 3.6 2.8<br />

CASP1 Casp1 caspase 1 12362 3.4 2.2<br />

IL-1β Il1b interleukin 1 beta 16176 29.0 21.5<br />

OAS Oas1g 2'-5' oligoadenylate synthetase 1G 23960 2.2 1.7<br />

Oas2 2'-5' oligoadenylate synthetase 2 246728 1.9 1.5<br />

Oas3 2'-5' oligoadenylate synthetase 3 246727 1.7 1.4<br />

RNaseL Rnasel ribonuclease L (2', 5'-oligoisoadenylate<br />

24014 1.6 1.4<br />

synthetase-dependent)<br />

transcription IRF-7 Irf7 interferon regulatory factor 7 54123 2.7 1.9<br />

factors (nucleus) NFκB Nfkb1 nuclear factor <strong>of</strong> kappa light polypeptide<br />

18033 2.0 1.9<br />

gene enhancer in B-cells 1, p105<br />

Nfkb2 nuclear factor <strong>of</strong> kappa light polypeptide<br />

18034 2.7 2.5<br />

gene enhancer in B-cells 2, p49/p100<br />

transcriptionally TNFα Tnf tumor necrosis factor 21926 3.0 2.3<br />

affected genes IL-6 Il6 interleukin 6 16193 9.2 7.8<br />

RANTES Ccl5 chemokine (C-C motif) ligand 5 20304 6.9 4.9<br />

a PRR – pattern recognition receptor<br />

Pattern recognition includes the complement system. The canonical pathway “Complement<br />

System” was significantly affected after infection (p = 6.86E−10; ratio = 16/36, Fig. R.2.8) which<br />

includes all three activation variants <strong>of</strong> classical, lectin and alternative pathway: The components<br />

C1q (C1qa, C1qb, C1qc), C1r (C1r, C1rb), C3, C4 (C4b), C7, Factor D / Adipsin (Cfd), Factor<br />

P / Properdin (Cfp), the complement-activating protease Masp1, and the receptors for C3a and<br />

C5a (C3ar, C5ar) were induced. A tendency <strong>of</strong> increased expression was observed for factors C2,<br />

C6, and Factor B (Cfb). But higher expression was also visible for the C1 inhibitor (Serping1), for<br />

the complement-inhibitory factor I (Cfi), and <strong>by</strong> trend for the inhibitory factor H (Cfh), probably as<br />

a reaction to restrict complement activation to a physiologically tolerable extent as the<br />

84


Maren Depke<br />

Results<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

complement system can cause immense damage also to <strong>host</strong> cells and tissue. The complement<br />

component C8 consists <strong>of</strong> three chains (alpha, beta, and gamma) and is the last component in the<br />

C5b-C6-C7-C8 complex that acts as a catalyst in the polymerization <strong>of</strong> C9 to form the membrane<br />

attack complex MAC. Surprisingly, the gene C8a coding for the alpha chain was repressed in the<br />

infected samples, and also for the gene C8g (gamma chain) a tendency <strong>of</strong> repression was<br />

recorded. Possibly the opsonization and chemoattractant function <strong>of</strong> the complement system<br />

was enhanced in infection leading to phagocytosis, while the bacteriolytic function <strong>of</strong> the<br />

complement system was not potentiated. Otherwise, the components C8 and C9 – produced in<br />

the liver – might be delivered via the blood stream to the site <strong>of</strong> infection.<br />

Fig. R.2.8:<br />

Complement System<br />

(modified from IPA, www.ingenuity.com).<br />

Colors indicate direction and magnitude <strong>of</strong><br />

differential regulation: red – induction; green –<br />

repression. More intense shade <strong>of</strong> color is used for<br />

higher absolute fold change values. A trend <strong>of</strong><br />

induction is shown in yellow. Factor P / Properdin<br />

(CFP) has been inserted manually into the pathway.<br />

As already mentioned, the Toll-like receptor (TLR) signaling cascades lead to the transcription<br />

<strong>of</strong> different proinflammatory cytokines which themselves activate signaling pathways and the<br />

adequate cellular responses. When using the Canonical Pathways to specifically find infection<br />

relevant signal transduction, the signaling pathways <strong>of</strong> interferon, IL-6, TREM1, and IL-10<br />

appeared among others with significant p-values.<br />

The pathway “Interferon Signaling” (p = 1.14E−06; ratio = 11/30; Fig. R.2.9) contains the<br />

induced IFN-γ receptor (Ifngr1, Ifngr2) and the IFNα/β-receptor, <strong>of</strong> which the gene for one <strong>of</strong> the<br />

two chains (Ifnar2) was induced, too. In the receptor’s signal transduction, the genes for signal<br />

transducers STAT 1 and 2 were induced together with the transcriptionally regulated genes Tap1,<br />

Ifitm1, Irf1, Psmb8, Irf9 and Oas1. Additionally, the inhibitory protein tyrosine phosphatase TC-<br />

PTP (Ptpn2) displayed a trend <strong>of</strong> induction, probably indicating a counter-regulatory process.<br />

85


Maren Depke<br />

Results<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

Fig. R.2.9:<br />

Interferon Signaling (modified<br />

from IPA, www.ingenuity.com).<br />

Red color indicates increase <strong>of</strong><br />

expression. More intense shade<br />

<strong>of</strong> color is used for higher<br />

absolute fold change values. A<br />

trend <strong>of</strong> induction is shown in<br />

yellow.<br />

Similarly, IL-6 receptor (Il6ra), IL-1 receptor (Il1r1, Il1r2) and TNF-α receptor (Tnfrsf1a,<br />

Tnfrsf1b) and their ligands (Il6, Il1a, Il1b, Tnf) were expressed at higher levels in infected tissue<br />

than in control tissue as depicted in the pathway “IL-6 Signaling” (p = 6.88E−11; ratio = 27/93;<br />

Fig. R.2.10). Glycoprotein 130 (Il6st), a signal transducer shared <strong>by</strong> IL-6 and other cytokines,<br />

exhibited at least a trend <strong>of</strong> induction. Again, genes for signal transduction molecules and<br />

transcription factors were significantly (Ikbke, Nfkbia, Nfkbia, Nfkbiz, Nfkb2, Stat3, Jun, Fos,<br />

Cebpb) or slightly (Nfkbid, Nfkb1, Traf2, Map4k4, Mapk13, Nras) increased. Signal transduction<br />

was activated in infected tissue, which was demonstrated <strong>by</strong> the induction <strong>of</strong> transcriptionally<br />

regulated genes like TSG6 (Tnfaip6), Collagen (Col1a1), alpha-2-macroglobulin (A2m), and IL-6<br />

itself.<br />

Fig. R.2.10: IL-6 Signaling (modified from IPA, www.ingenuity.com).<br />

Direction and magnitude <strong>of</strong> differential expression are indicated <strong>by</strong> colors: red – induction; green – repression. More intense shade <strong>of</strong><br />

color is used for higher absolute fold change values. A trend <strong>of</strong> induction is shown in yellow.<br />

86


Maren Depke<br />

Results<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

The canonical pathway “TREM1 Signaling” (p = 8.15E−13; ratio = 23/69) slightly overlaps with<br />

the pathway “IL-6 Signaling” as branches <strong>of</strong> signal transduction also lead to IL-6 and TNFα<br />

transcription via STAT3 and NFκB (all induced after infection). Also TLR4 is included because after<br />

activation, the receptor can associate with TREM1, activate the kinase IRAK1 and finally the<br />

NFκB-mediated transcription. The increased Trem1 itself, a receptor whose ligand is still<br />

unknown and which is expressed <strong>by</strong> neutrophils, monocytes and macrophages, and the increased<br />

adapter molecule DAP12 (Tyrobp) are the beginning <strong>of</strong> a proinflammatory reaction chain that<br />

includes AKT/PKB (Rac2, induced), PLCγ and NTAL (Plcg2 and Lat2, induced <strong>by</strong> trend). TREM1<br />

signaling not only influences transcription <strong>of</strong> proinflammatory cytokines (MCP-1/Ccl2, TNFα,<br />

MCP-3/Ccl7, IL-6, MIP-1α/Ccl3), but also the translocation <strong>of</strong> cytokines from cytoplasm to the<br />

extracellular space (MCP-1/Ccl2, TNFα, IL-6, IL-1β). DAP12 positively influences the activation <strong>of</strong><br />

IL-1β <strong>by</strong> CASP1 (both induced) in the pathway <strong>of</strong> NLR (intracellular NACHT-LRR receptors)<br />

recognizing intracellular bacteria.<br />

Furthermore, DAP12 leads to the transcription <strong>of</strong> cell adhesion molecules (CD11c/Itgax,<br />

CD49e/Itga5) and the co-stimulatory proteins CD54/Icam1 and CD86 (all induced). Genes coding<br />

for cell surface proteins CD32/Fcgr2b and CD40, which are additionally regulated via DAP12, were<br />

increased <strong>by</strong> trend.<br />

Opposed to these proinflammatory signaling pathways, the “IL-10 Signaling” pathway<br />

(p = 1.99E−15; ratio = 27/70; Fig. R.2.11) is an example for a mechanism aiming to limit and<br />

terminate the cellular inflammatory response in addition to regulating growth and differentiation<br />

<strong>of</strong> different immune cells. Here, the IL-10 receptor α-chain (Il10ra) was induced after infection.<br />

Induced Stat3 is part <strong>of</strong> the signaling cascade like in IL-6 signaling but SOCS3, suppressor <strong>of</strong><br />

cytokine signaling, was additionally increased. This gene is itself transcriptionally regulated at the<br />

end <strong>of</strong> the IL-10 signaling pathway. Furthermore, Ccr1, Ccr5, Arg2, Il4ra, and Hmox1 were<br />

induced genes responding to IL-10. IL-6 signals are also mediated via STAT3, and the cytokine IL-6<br />

was induced contrarily to IL-10, which was not regulated. Therefore, the IL-10 pathway might not<br />

be activated yet, but just prepared for a later phase <strong>of</strong> anti-inflammatory reactions.<br />

Fig. R.2.11:<br />

IL-10 Signaling (modified from IPA, www.ingenuity.com).<br />

Red color indicates increase <strong>of</strong> expression. More intense shade <strong>of</strong> color is<br />

used for higher absolute fold change values.<br />

87


Maren Depke<br />

Results<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

An essential part <strong>of</strong> the immune response to infection is the presentation <strong>of</strong> antigens to<br />

immune cells to direct the defense against the specific <strong>pathogen</strong> in the adaptive part <strong>of</strong> the<br />

immune system. Antigens belong to two main groups, extracellular and intracellular antigens,<br />

which are processed in separate pathways. Extracellular antigens are taken up <strong>by</strong><br />

phagocytosis/endocytosis and digested in vesicles after fusion with lysosomes. Finally, peptides<br />

are bound to MHC-II molecules originating from exocytic Golgi vesicles and presented on the cell<br />

surface. Intracellular antigens are digested <strong>by</strong> cytosolic proteasome with sequential transport <strong>of</strong><br />

peptides into the endoplasmatic reticulum (ER). These peptides are loaded to MHC-I molecules<br />

and transported via the Golgi to the cell surface for presentation. Presentation <strong>of</strong> extracellular<br />

antigens preferentially occurs <strong>by</strong> pr<strong>of</strong>essional antigen presenting cells (APC) like dendritic cells<br />

(DC), macrophages and B cells. The peptide-MHC-II complex is recognized <strong>by</strong> CD4 + helper T cells,<br />

which leads to initiation <strong>of</strong> mechanisms fighting extracellular infection or disposing extracellular<br />

antigens. Immune cells and other cells <strong>of</strong> the body are capable <strong>of</strong> presenting intracellular<br />

antigens via MHC-I. They present the antigenic peptide to CD8 + cytotoxic T cells and there<strong>by</strong><br />

activate processes to destroy cells and kill <strong>pathogen</strong> in case <strong>of</strong> intracellular infection.<br />

For both pathways an increase in transcription <strong>of</strong> associated genes was observed after<br />

infection (Fig. R.2.12). The immune-proteasome genes Psmb9, Psmb10, and Psmb8 (LMP2,<br />

LMP7), coding for proteins in the 20S-core <strong>of</strong> the proteasome, were induced leading to an<br />

immune-response specific reorganization <strong>of</strong> the proteasomes’ catalytic center. For the genes<br />

Psme1 and Psme2, which contribute to the 11S regulatory complex <strong>of</strong> the proteasome, a trend <strong>of</strong><br />

induction was visible. Induced were also the peptide transporters Tap1 and Tap2 and the TAPbinding<br />

protein tapasin/TPN (Tapbp), which attaches the TAP molecules to the MHC-I complex.<br />

The ER aminopeptidase Erap1, which is responsible for N-terminal trimming <strong>of</strong> the peptide in the<br />

ER, was induced in tendency. Finally, MHC-I molecules (H2-D1, H2-K1, H2-Q6, H2-Q7, H2-Q8) and<br />

Beta-2-microglobulin (MHC-I-β; B2m) were induced, too.<br />

In the pathway for presentation <strong>of</strong> extracellular antigens MHC-II molecules (H2-Aa, H2-Ab1,<br />

H2-DMa, H2-DMb1, H2-Ea, H2-Eb1) and the invariant chain Cd74 (CLIP) were increased. Induction<br />

<strong>of</strong> lysosomal enzymes, particularly <strong>of</strong> cathepsins, was also observed (Chi3l1, Chi3l3, Hpse, Ctsc,<br />

Ctse, Ctss; induced <strong>by</strong> trend: Gla, Ctsd, Ctsk, Ctsl).<br />

Fig. R.2.12: Antigen Presentation (modified from IPA, www.ingenuity.com). Red color indicates increase <strong>of</strong> expression.<br />

88


Maren Depke<br />

Results<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

Metabolic effects <strong>of</strong> infection in kidney tissue<br />

In an approach to recognize changes in gene expression <strong>of</strong> metabolic enzymes after infection<br />

lists <strong>of</strong> differentially regulated genes were subjected to a BIOCYC “omics-Viewer” analysis (BIOCYC,<br />

SRI International, CA, USA, http://biocyc.org/expression.html). This tool allows the display <strong>of</strong><br />

gene expression data on highly abstracted metabolic pathway schemes and therefore an intuitive<br />

comprehension <strong>of</strong> processes in the selected experimental setup. Two different lists were<br />

included in the analysis: a) genes significantly different between infection and sham infection<br />

with a minimal absolute fold change <strong>of</strong> 2 in at least one <strong>of</strong> the two comparisons “infection with<br />

RN1HG vs. sham infection” and “infection with RN1HG ΔsigB vs. sham infection” and b) genes<br />

significantly different between infection and sham infection with a minimal absolute fold change<br />

<strong>of</strong> 1.5 in at least one <strong>of</strong> the two comparisons described before.<br />

When focusing on regulation that exceeds a factor <strong>of</strong> 2 in at least one comparison, regulation<br />

was discernible for certain groups <strong>of</strong> metabolic pathways (Fig. R.2.13 A). After infection, a<br />

repression <strong>of</strong> genes relevant for cholesterol biosynthesis was visible, and also pathways <strong>of</strong> amino<br />

acid degradation contained repressed genes. Induction <strong>of</strong> gene expression was detected for<br />

steroid hormone biosynthesis and for purine degradation. This is surprising, as induction was also<br />

visible for steps <strong>of</strong> the purine/pyrimidine biosynthesis. Similarly, an increase in gene expression<br />

was measured for certain amino acid biosynthesis steps.<br />

Including differentially expressed genes with lower absolute fold change values (1.5 in at least<br />

one <strong>of</strong> two comparisons) supplemented the data set with many repressed and only few induced<br />

genes, which is visualized <strong>by</strong> the over-representation <strong>of</strong> yellow colored reaction steps<br />

(Fig. R.2.13 B). In this view, repression in further steps in the group <strong>of</strong> lipid biosynthesis and<br />

amino acid degradation was visible. Additional repression occurred in aerobic respiration, TCA<br />

cycle, glycolysis, but also in gluconeogenesis and glycogen biosynthesis. The pattern <strong>of</strong> increased<br />

genes dominating both purine degradation as well as biosynthesis was maintained after inclusion<br />

<strong>of</strong> regulation with lower absolute fold change values.<br />

As the gene expression analysis was performed for tissue samples, i. e. <strong>of</strong> a mixture <strong>of</strong><br />

different kidney tissue cell types and in case <strong>of</strong> infection additionally <strong>of</strong> a mixture <strong>of</strong> different<br />

immune cell types invading the inflamed tissue, the resulting pattern is difficult to match for<br />

particular cell types. While one cell type might induce purine degradation e. g. for using it as<br />

catabolic substrate, another cell type might induce purine biosynthesis e. g. as basis for DNA<br />

replication during proliferation. Such situation could lead to contradictory results concerning<br />

metabolic pathways if the number <strong>of</strong> cells influencing the pattern in a certain direction is high<br />

enough.<br />

The general gene expression pattern showed a metabolic disturbance in several pathways<br />

that cannot be distinguished into anabolic or catabolic shift because it contained reduction <strong>of</strong><br />

several catabolic pathways like glycolysis, amino acid degradation, aerobic respiration, and TCA<br />

cycle, but also reduction in anabolic reactions like gluconeogenesis or amino acid biosynthesis.<br />

The reason <strong>of</strong> this unclear or mixed reaction might be found in the high extent <strong>of</strong> infection and<br />

illness. Possibly the infection has caused such a strong damage to the kidney after 5 days that<br />

also organ function, nutrition supply and possibly oxygen availability are impaired.<br />

89


Maren Depke<br />

Results<br />

Kidney Gene Expression Pattern in an in vivo Infection Model<br />

A<br />

co-factor biosynthesis<br />

cholesterol,<br />

triacylglycerol,<br />

phospholipid<br />

biosynthesis<br />

aerobic<br />

respiration<br />

amino acid degradation<br />

amino acid<br />

biosynthesis<br />

other<br />

degradation<br />

pathways<br />

purine/<br />

pyrimidine<br />

biosynthesis<br />

sugar<br />

degradation<br />

gluconeogenesis steroid<br />

and glycogen hormone<br />

biosynthesis biosynthesis<br />

glycolysis<br />

prostaglandin<br />

biosynthesis<br />

TCA<br />

purine<br />

degradation<br />

fatty acid oxidation,<br />

triacylglycerol,<br />

phospholipid<br />

degradation<br />

B<br />

co-factor biosynthesis<br />

cholesterol,<br />

triacylglycerol,<br />

phospholipid<br />

biosynthesis<br />

aerobic<br />

respiration<br />

amino acid degradation<br />

amino acid<br />

biosynthesis<br />

other<br />

degradation<br />

pathways<br />

purine/<br />

pyrimidine<br />

biosynthesis<br />

sugar<br />

degradation<br />

gluconeogenesis steroid<br />

and glycogen hormone<br />

biosynthesis biosynthesis<br />

glycolysis<br />

prostaglandin<br />

biosynthesis<br />

TCA<br />

purine<br />

degradation<br />

fatty acid oxidation,<br />

triacylglycerol,<br />

phospholipid<br />

degradation<br />

Fig. R.2.13: The influence <strong>of</strong> infection on gene expression in murine metabolic pathways (modified from omics-viewer(s) <strong>of</strong> BIOCYC,<br />

SRI International, CA, USA, http://biocyc.org/expression.html).<br />

Nodes represent metabolites and lines indicate reactions. The metabolic reactions are colored according to the enzymes’ gene<br />

expression regulation in the comparison <strong>of</strong> infection vs. sham infection. Red marks an increase and yellow a decrease in infected<br />

tissue. The display is limited to genes whose absolute fold change exceeds 2 in at least one <strong>of</strong> the two comparisons “infection with<br />

RN1HG vs. sham infection” and “infection with RN1HG ΔsigB vs. sham infection” (A) or to those whose absolute fold change exceeds<br />

1.5 in at least one <strong>of</strong> the two comparisons mentioned before (B).<br />

90


Maren Depke<br />

Results<br />

GENE EXPRESSION PATTERN OF BONE-MARROW DERIVED<br />

MACROPHAGES AFTER INTERFERON-GAMMA TREATMENT<br />

Experimental application <strong>of</strong> serum-free differentiation and cultivation <strong>of</strong> mouse BMM<br />

Bone-marrow derived macrophages (BMM) allow experimental analysis <strong>of</strong> macrophage<br />

reactions and functions without the impact <strong>of</strong> immunological conditioning that affects already<br />

differentiated macrophages for example from spleen or peritoneum. The standardization <strong>of</strong><br />

experiments was markedly improved in 2009 when Eske et al. published a serum-free culture<br />

system for BMM that allows conduction <strong>of</strong> studies independently from the influence <strong>of</strong><br />

undefined immune-active substances and possible variation introduced to experiments <strong>by</strong> using<br />

fetal calf serum (FCS) as supplement. BMM cultivated in this new serum-free medium maintain<br />

the already known characteristics <strong>of</strong> macrophages and have proven to lead to highly reproducible<br />

results (Eske et al. 2009).<br />

The experimental setup in this study included the two mouse strains BALB/c and C57BL/6.<br />

After in vitro differentiation, BMM <strong>of</strong> each strain were divided in two treatment groups: IFN-γ<br />

treated BMM and non-treated medium control BMM.<br />

The approach <strong>of</strong> transcriptome analysis using the Affymetrix GeneChip Mouse Gene 1.0 ST<br />

allowed monitoring <strong>of</strong> 35557 probe sets, <strong>of</strong> which 6613 belong to the group <strong>of</strong> controls. The<br />

remaining 28944 probe sets represent 20074 EntrezGene records in the annotations <strong>of</strong> Rosetta<br />

Resolver s<strong>of</strong>tware (annotation <strong>of</strong> 06/2009). The LC-MS/MS proteome analysis provided<br />

information on 946 reliably identified proteins (Dinh Hoang Dang Khoa).<br />

High reproducibility <strong>of</strong> gene expression in experiments using serum-free differentiation and<br />

cultivation <strong>of</strong> mouse BMM<br />

For a first general impression <strong>of</strong> the array data set, the method <strong>of</strong> Principal Component<br />

Analysis (PCA) was applied. This method calculates the direction <strong>of</strong> strongest variation from the<br />

multidimensional array data set, and reduces it to a new value <strong>of</strong> the parameter called Principle<br />

Component (PC). The remaining variation in the data set is subsequently addressed in the same<br />

way until all or a pre-defined fraction <strong>of</strong> variation is collapsed into new values. This procedure<br />

results in a set <strong>of</strong> PCs, <strong>of</strong> which each accounts for a fraction <strong>of</strong> the total variance in the data set.<br />

Usually, the first 2 or 3 PCs are displayed in a 2- or 3-dimensional coordinate system, respectively.<br />

In such a plot, the distance <strong>of</strong> the points that represent the individual data sets correlates to the<br />

difference between them.<br />

In this study, the PCA plot was derived from log-transformed intensity data <strong>of</strong> 12 arrays,<br />

analyzing 3 biological replicates <strong>of</strong> 2 strains and 2 treatment groups (Fig. R.3.1). The PCA clearly<br />

depicted the high reproducibility <strong>of</strong> biological replicates obtained from the serum-free model <strong>of</strong><br />

BMM differentiation, cultivation and IFN-γ treatment: The 3 biological replicates <strong>of</strong> each group<br />

were arranged together, while the BMM <strong>of</strong> different strains and treatments were separated.<br />

Differences on the axis for PC 1 were <strong>by</strong> a factor <strong>of</strong> approximately 250 smaller than differences<br />

on the axes for PCs 2 and 3. That implied that PCs 2 and 3 covered the main part <strong>of</strong> variance in<br />

the data set. This led to the conclusion that the data set was mainly influenced <strong>by</strong> 2 factors. In an<br />

experimental design <strong>of</strong> the 2 factors strain and treatment together with the observed grouping <strong>of</strong><br />

biological replicates obviously the experimental factors were responsible for the observed<br />

variation.<br />

91


Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

Fig. R.3.1:<br />

Transcriptome data visualization using Principal<br />

Component Analysis (PCA).<br />

Log-intensity data were used to derive a PCA<br />

plot containing 12 arrays (analyzing 3 biological<br />

replicates <strong>of</strong> 2 strains and 2 treatment groups).<br />

Resulting principal components were set to<br />

cover 95 % <strong>of</strong> total variation, while the total<br />

number <strong>of</strong> principal components was not<br />

limited. The analysis was restricted to noncontrol<br />

probe sets that were not absent on all<br />

12 arrays, when absence <strong>of</strong> expression is<br />

defined via a p-value > 0.01 on pr<strong>of</strong>ile level in<br />

Rosetta Resolver. Differences on the axis for<br />

principal component 1 are <strong>by</strong> a factor <strong>of</strong><br />

approximately 250 smaller than differences on<br />

the axis for principal components 2 and 3,<br />

which is shown in the small insert that depicts<br />

the same PCA but is turned <strong>by</strong> 90° to the right<br />

around the axis <strong>of</strong> principle component 2.<br />

Arrays <strong>of</strong> BALB/c BMM samples are<br />

represented <strong>by</strong> dots (•), arrays <strong>of</strong> C57BL/6<br />

BMM samples <strong>by</strong> rhombi (). IFN-γ treated<br />

samples are indicated in gray, and non-treated<br />

control level samples are colored in black.<br />

As transcriptome data demonstrated a high reproducibility <strong>of</strong> biological replicates in the<br />

model system, selected samples were used for different proteome applications rather than<br />

analyzing different biological replicates with only one proteomics approach (Dinh Hoang Dang<br />

Khoa).<br />

For a general comparison, gelfree proteome analysis and transcriptome analysis are more<br />

comparable than 2-DE proteome analysis and transcriptome analysis. Therefore, further data<br />

examination focused on that.<br />

Overall comparison <strong>of</strong> transcriptome [Maren Depke] and LC-MS/MS proteome [Dinh Hoang<br />

Dang Khoa] analyses<br />

900 <strong>of</strong> 946 identified proteins from gel-free LC-MS/MS approach could be mapped to the<br />

20074 EntrezGene records that were specified to be represented on the Affymetrix GeneChip<br />

Mouse Gene 1.0 ST array <strong>by</strong> the Rosetta Resolver s<strong>of</strong>tware (Table R.3.1). As expected, a much<br />

bigger number <strong>of</strong> genes was accessible <strong>by</strong> transcriptome analysis than proteins <strong>by</strong> gelfree<br />

proteome analysis. Nevertheless, for almost all identified proteins the gene expression pattern<br />

could be recorded.<br />

Table R.3.1: Overall comparison transcriptome [Maren Depke] and LC-MS/MS proteome [Dinh Hoang Dang Khoa] analyses and<br />

numbers <strong>of</strong> differentially regulated genes and proteins with differing abundance resulting from LC-MS/MS analysis.<br />

total number <strong>of</strong><br />

detected genes or<br />

proteins<br />

in BALB/c BMM<br />

IFN-γ effects a<br />

in C57BL/6 BMM<br />

at non-treated<br />

control level<br />

strain differences a<br />

at IFN-γ<br />

treated level<br />

microarray 20074 442 (2.2 %) 396 (2.0 %) 222 (1.1 %) 230 (1.1 %)<br />

LC-MS/MS 946 45 (4.8 %) 52 (5.5 %) 218 (23.0 %) 308 (32.6 %)<br />

a Percentage values indicate the proportion <strong>of</strong> regulation in both approaches. Values were calculated <strong>by</strong> dividing the number <strong>of</strong><br />

regulated genes <strong>by</strong> the total number <strong>of</strong> genes available on the array or the number <strong>of</strong> regulated proteins <strong>by</strong> the total number <strong>of</strong><br />

identified proteins.<br />

92


Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

Surprisingly, the percentage <strong>of</strong> proteins with significant difference in abundance between<br />

BMM <strong>of</strong> BALB/c and C57BL/6 mice compared to the total <strong>of</strong> all identified proteins was noticeably<br />

higher than the percentage <strong>of</strong> differentially expressed genes between BMM <strong>of</strong> BALB/c and<br />

C57BL/6 mice compared to the total <strong>of</strong> all genes available on the array (Table R.3.1).<br />

IFN-γ influence on gene expression [Maren Depke] and protein abundances [Dinh Hoang Dang<br />

Khoa] in BMM <strong>of</strong> BALB/c and C57BL/6 mice<br />

The comparison <strong>of</strong> IFN-γ treated BMM with the non-treated medium control revealed that<br />

IFN-γ mainly led to induction <strong>of</strong> gene expression or increase in protein abundance in both strains<br />

(Table R.3.2 A). From the total <strong>of</strong> 442 (BALB/c BMM) and 396 (C57BL/6 BMM) differentially<br />

expressed genes, 388 (BALB/c BMM) and 330 (C57BL/6 BMM) were induced, while only 54<br />

(BALB/c BMM) and 66 (C57BL/6 BMM) were repressed.<br />

IFN-γ treatment increased the abundance <strong>of</strong> 41 (BALB/c BMM) and 43 (C57BL/6 BMM)<br />

proteins and reduced the abundance <strong>of</strong> 4 (BALB/c BMM) and 9 (C57BL/6 BMM) from the total <strong>of</strong><br />

45 (BALB/c BMM) and 52 (C57BL/6 BMM) regulated proteins.<br />

The overlap <strong>of</strong> the list <strong>of</strong> proteins that differed significantly in their abundance between IFN-γ<br />

treated BMM and the non-treated medium control and the list <strong>of</strong> differentially expressed genes<br />

that were found after IFN-γ treatment amounted to 24 (BALB/c BMM) and 20 (C57BL/6 BMM).<br />

Compared to 43 (BALB/c BMM) and 50 (C57BL/6 BMM) regulated proteins also available <strong>by</strong><br />

transcriptome analysis and 52 (BALB/c BMM) and 53 (C57BL/6 BMM) differentially expressed<br />

genes also accessible <strong>by</strong> LC-MS/MS proteome analysis, the overlap <strong>of</strong> regulated proteins and<br />

genes added up to 38 % to 56 % (Table R.3.2 A).<br />

Table R.3.2: Comparison <strong>of</strong> differentially regulated genes resulting from transcriptome analysis [Maren Depke] with differentially<br />

regulated proteins resulting from LC-MS/MS analysis [Dinh Hoang Dang Khoa].<br />

A Direction <strong>of</strong> regulation <strong>by</strong> IFN-γ in BALB/c BMM and<br />

in C57BL/6 BMM.<br />

IFN-γ treatment effects<br />

B Direction <strong>of</strong> strain difference between C57BL/6 BMM and BALB/c<br />

BMM at non-treated medium-control and IFN-γ treated level.<br />

strain differences<br />

LC-<br />

a intersection<br />

a microarray<br />

MS/MS<br />

LC-<br />

a intersection<br />

a microarray<br />

MS/MS<br />

a<br />

BALB/c BMM<br />

control condition BMM<br />

induced 39 (41) 41 (388) 24 higher in C57BL/6 102 (112) 10 (100) 7<br />

repressed 4 (4) 11 (54) 0 higher in BALB/c 101 (106) 10 (122) 4<br />

total 43 (45) 52 (442) 24 total 203 (218) 20 (222) 12<br />

C57BL/6 BMM<br />

treatment condition BMM<br />

induced 42 (43) 41 (330) 20 higher in C57BL/6 +IFN-γ 123 (135) 13 (93) 10<br />

repressed 8 (9) 12 (66) 0 higher in BALB/c +IFN-γ 165 (173) 11 (137) 7<br />

total 50 (52) 53 (396) 20 Total 288 (308) 24 (230) 19<br />

The first number in the table always refers to the proteins or genes that were available for analysis with both approaches,<br />

microarray and LC-MS/MS analysis. In brackets, the number <strong>of</strong> regulated proteins or genes without restriction to the accessibility <strong>by</strong><br />

both methods is given.<br />

The comparison <strong>of</strong> IFN-γ effects in BMM <strong>of</strong> the two strains BALB/c and C57BL/6 resulted in a<br />

high overlap <strong>of</strong> 307 differentially expressed genes (Fig. R.3.2 A) and 29 proteins with significantly<br />

different abundance (Fig. R.3.2 C). Even when the gene or protein was not significantly expressed<br />

in one <strong>of</strong> the strains a similar expression trend in BMM <strong>of</strong> both strains was observed (Fig. R.3.2 B,<br />

D).<br />

93


log2(ratio IFN-γ-treated / control) <strong>of</strong> C57BL/6 BMM<br />

log2(ratio IFN-γ-treated / control) <strong>of</strong> C57BL/6 BMM<br />

Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

A<br />

B<br />

10<br />

8<br />

6<br />

4<br />

135 307 89<br />

2<br />

0<br />

-2<br />

genes regulated <strong>by</strong> IFN-γ<br />

in BALB/c BMM (442)<br />

genes regulated <strong>by</strong> IFN-γ<br />

in C57BL/6 BMM (396)<br />

-4<br />

-6<br />

-8<br />

-10<br />

-10 -8 -6 -4 -2 0 2 4 6 8 10<br />

log 2 (ratio IFN-γ-treated / control) <strong>of</strong> BALB/c BMM<br />

C<br />

D<br />

10<br />

8<br />

6<br />

4<br />

16 29 23<br />

2<br />

0<br />

-2<br />

proteins regulated <strong>by</strong> IFN-γ<br />

in BALB/c BMM (45)<br />

proteins regulated <strong>by</strong> IFN-γ<br />

in C57BL/6 BMM (52)<br />

-4<br />

-6<br />

-8<br />

-10<br />

-10 -8 -6 -4 -2 0 2 4 6 8 10<br />

log 2 (ratio IFN-γ-treated / control) <strong>of</strong> BALB/c BMM<br />

Fig. R.3.2: Comparison <strong>of</strong> IFN-γ effects on gene expression pattern and protein abundance in BALB/c BMM and C57BL/6 BMM.<br />

A, C. Overview on numbers <strong>of</strong> differentially expressed genes in transcriptome analysis (A) and <strong>of</strong> proteins with significantly different<br />

abundance in LC-MS/MS analyses (C) when comparing IFN-γ effects in BALB/c BMM and C57BL/6 BMM.<br />

B, D. Log 2-transformed ratio data “IFN-γ treated BALB/c BMM” / “medium control BALB/c BMM” were plotted on the x-axis and log 2-<br />

transformed ratio data “IFN-γ treated C57BL/6 BMM” / “medium control C57BL/6 BMM” on the y-axis from transcriptome (B) and LC-<br />

MS/MS (D) analyses.<br />

Coloring distinguishes three different groups <strong>of</strong> regulated genes or proteins: Genes/proteins with significant differential<br />

expression/abundance caused <strong>by</strong> IFN-γ only in BALB/c BMM are shown in dark gray, while that with significant differential<br />

expression/abundance caused <strong>by</strong> IFN-γ only in C57BL/6 BMM are displayed in light gray. Genes/proteins significantly regulated <strong>by</strong><br />

IFN-γ in both BALB/c BMM and C57BL/6 BMM are colored in black. The minimal absolute fold change cut<strong>of</strong>f <strong>of</strong> 1.5 has been applied to<br />

all result lists.<br />

All transcriptome ratio data were derived from mean intensities <strong>of</strong> three biological replicates, and ratio data <strong>of</strong> LC-MS/MS analyses<br />

were derived from intensities <strong>of</strong> one biological replicate analyzed in technical triplicates <strong>of</strong> each group.<br />

Strain differences in gene expression [Maren Depke] and protein abundances [Dinh Hoang<br />

Dang Khoa] between BMM <strong>of</strong> BALB/c and C57BL/6 mice in untreated medium control and in<br />

IFN-γ treated samples<br />

The comparison <strong>of</strong> C57BL/6 BMM and BALB/c BMM at the non-treated medium control level<br />

and at the IFN-γ treated level resulted in 100 (control level) and 93 (IFN-γ treated level) genes<br />

that featured a higher expression in C57BL/6 BMM and 122 (control level) and 137 (IFN-γ treated<br />

level) genes that exhibited a higher expression in BALB/c BMM.<br />

94


log2(ratio C57BL/6 / BALB/c) at IFN-γ-activated level<br />

log2(ratio C57BL/6 / BALB/c) at IFN-γ-activated level<br />

Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

The same comparison using the LC-MS/MS proteome data yielded 112 (control level) and 135<br />

(IFN-γ treated level) proteins with a higher abundance in C57BL/6 BMM and 106 (control level)<br />

and 173 (IFN-γ treated level) proteins with a higher abundance in BALB/c BMM (Table R.3.2 B).<br />

A<br />

B<br />

10<br />

8<br />

6<br />

94 128 102<br />

4<br />

2<br />

genes differentially<br />

expressed between<br />

BALB/c BMM and C57BL/6<br />

BMM at the non-activated<br />

control level (222)<br />

genes differentially<br />

expressed between<br />

BALB/c BMM and C57BL/6<br />

BMM at the IFN-γ-treated<br />

level (230)<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

-10<br />

-10 -8 -6 -4 -2 0 2 4 6 8 10<br />

log 2 (ratio C57BL/6 / BALB/c) at control level<br />

C<br />

D<br />

10<br />

8<br />

6<br />

35 183 125<br />

4<br />

2<br />

proteins differentially<br />

regulated between BALB/c<br />

BMM and C57BL/6 BMM<br />

at the non-activated<br />

control level (218)<br />

proteins differentially<br />

regulated between BALB/c<br />

BMM and C57BL/6 BMM<br />

at the IFN-γ-treated level<br />

(308)<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

-10<br />

-10 -8 -6 -4 -2 0 2 4 6 8 10<br />

log 2 (ratio C57BL/6 / BALB/c) at control level<br />

Fig. R.3.3: Comparison <strong>of</strong> strain differences between BALB/c BMM and C57BL/6 BMM at non-treated control and IFN-γ treated level.<br />

A, C. Overview on numbers <strong>of</strong> differentially expressed genes in transcriptome analysis (A) and <strong>of</strong> proteins with significantly different<br />

abundance in LC-MS/MS analyses (C) when comparing strain differences between BALB/c BMM and C57BL/6 BMM at the non-treated<br />

control level and at the IFN-γ treated level.<br />

B, D. Log 2-transformed ratio data “medium control C57BL/6 BMM” / “medium control BALB/c BMM” were plotted on the x-axis and<br />

log 2-transformed ratio data “IFN-γ treated C57BL/6 BMM” / “IFN-γ treated BALB/c BMM” on the y-axis from transcriptome (B) and LC-<br />

MS/MS (D) analyses.<br />

Coloring distinguishes three different groups <strong>of</strong> regulated genes or proteins: Genes/proteins with significant strain difference only at<br />

non-treated control level are shown in dark gray, while that with significant strain difference only at IFN-γ treated level are displayed<br />

in light gray. Genes/proteins significantly different between BALB/c BMM and C57BL/6 BMM on both treatment levels, i. e. at nontreated<br />

control and IFN-γ treated level, are colored in black. The minimal absolute fold change cut<strong>of</strong>f <strong>of</strong> 1.5 has been applied to all<br />

result lists.<br />

All transcriptome ratio data were derived from mean intensities <strong>of</strong> three biological replicates, and ratio data <strong>of</strong> LC-MS/MS analyses<br />

were derived from intensities <strong>of</strong> one biological replicate analyzed in technical triplicates <strong>of</strong> each group.<br />

In the comparison <strong>of</strong> the two mouse strains only 20 out <strong>of</strong> 222 (control level) and 24 out <strong>of</strong><br />

230 (IFN-γ treated level) differentially expressed genes were also identified in LC-MS/MS analysis.<br />

Inversely, 203 out <strong>of</strong> 218 (control level) and 288 out <strong>of</strong> 308 (IFN-γ treated level) proteins with<br />

significantly differing abundance were also detected on the Affymetrix array. Due to the limiting<br />

95


Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

number <strong>of</strong> genes whose corresponding proteins were accessible <strong>by</strong> LC-MS/MS analysis, the<br />

overlap <strong>of</strong> differentially abundant proteins and differentially expressed genes was low in absolute<br />

number (12 at control level and 19 at IFN-γ treated level) but it still held 60 % (control level) and<br />

79 % (IFN-γ treated level) <strong>of</strong> the genes that were both differentially expressed in the array data<br />

set and accessible <strong>by</strong> proteome analysis (Table R.3.2 B). Vice versa, for only 5.9 % (control level)<br />

and 6.6 % (IFN-γ treated level) <strong>of</strong> regulated proteins with accessible microarray data a change in<br />

gene expression level was detected. Hence, the majority <strong>of</strong> protein abundance differences<br />

between the strains were independent from gene expression level.<br />

A refined comparison between proteome and transcriptome analysis which included the<br />

direction <strong>of</strong> regulation uncovered that the overlap <strong>of</strong> proteins and genes higher expressed in<br />

C57BL/6 BMM and the overlap <strong>of</strong> proteins and genes higher expressed in BALB/c BMM (7 and 4<br />

at control level; 10 and 17 at IFN-γ treated level) did not sum up to the total overlap <strong>of</strong> 12 at<br />

control level and 19 at IFN-γ treated level (Table R.3.2 B). The explanation for the discrepancy is<br />

that one protein/gene (Scp2) at the control level showed a higher expression in BALB/c BMM in<br />

transcriptome but higher abundance in C57BL/6 BMM in proteome analysis. Correspondingly, at<br />

the IFN-γ treated level two proteins/genes did not possess the same direction <strong>of</strong> regulation: H2-<br />

AB1 and H2-K1 were higher expressed in C57BL/6 BMM in transcriptome but showed higher<br />

abundance in BALB/c BMM in proteome analysis.<br />

The comparison <strong>of</strong> strain differences which occurred at the non-treated medium control level<br />

and at the IFN-γ treated level demonstrated similar differences between C57BL/6 BMM and<br />

BALB/c BMM at both treatment levels: 128 genes were expressed differentially at both treatment<br />

conditions (Fig. R.3.3 A), and 183 proteins show significantly different abundance (Fig. R.3.3 C).<br />

Equivalent to the IFN-γ effects, a similar expression trend in the strain comparisons was visible at<br />

both treatment levels, even when regulation was significant only in one (Fig. R.3.3 B, D).<br />

Revelation <strong>of</strong> similar main effects <strong>of</strong> IFN-γ in BALB/c BMM and C57BL/6 BMM <strong>by</strong> network<br />

analysis and confirmation <strong>of</strong> known IFN-γ effects in serum-free BMM cultivation system<br />

Ingenuity Pathway Analysis (IPA, www.ingenuity.com) allows to arrange genes <strong>of</strong> interest, e. g.<br />

differentially expressed genes, in networks which show the <strong>interactions</strong> in the selected group <strong>of</strong><br />

genes. The two networks with the highest score from the separate analyses for IFN-γ effects in<br />

BALB/c BMM and C57BL/6 BMM exhibited an overlap <strong>of</strong> 22 out <strong>of</strong> 35 genes (nodes). Therefore,<br />

they covered similar functional aspects. After merging both networks, the comparison <strong>of</strong><br />

differential gene expression after IFN-γ treatment in BALB/c BMM (Fig. R.3.4 A) and in C57BL/6<br />

BMM (Fig. R.3.4 B) showed that many <strong>of</strong> the genes <strong>of</strong> this network were subjected to comparable<br />

regulation <strong>by</strong> IFN-γ in BMM <strong>of</strong> both strains, sometimes with differences in magnitude.<br />

The network confirms already known IFN-γ effects in the new serum-free BMM cultivation<br />

system. The 26 S proteasome was also known to be changed in response to IFN-γ stimulus. The<br />

proteasome, consisting <strong>of</strong> a ring-shaped 20 S core complex <strong>of</strong> several subunits, which exhibits<br />

besides others protease function, and two regulatory complexes <strong>of</strong> 19 S, is the center <strong>of</strong><br />

intracellular ubiquitin-dependent protein degradation which is not only needed for the normal<br />

protein turnover but also for the generation <strong>of</strong> peptides for the presentation to immune cells via<br />

MHC-I molecules. The change <strong>of</strong> the normal proteasome into the so-called immunoproteasome is<br />

mediated <strong>by</strong> the exchange <strong>of</strong> three β-subunits from the core complex <strong>by</strong> gene products <strong>of</strong><br />

Psmb9, Psmb10, and Psmb8 and is induced <strong>by</strong> IFN-γ. The exchange <strong>of</strong> these subunits leads to a<br />

change in the proteolytic activity and specificity (Wang/Maldonado 2006, Strehl et al. 2005). Also<br />

an alternative regulatory complex <strong>of</strong> 11 S (PA28) exists (Rivett et al. 2001). Here, expression <strong>of</strong><br />

Psmb9 (3.6 / 3.4), Psmb10 (3.1 / 3.7), and Psmb8 (2.8 / 3.1), coding for β-subunits <strong>of</strong> the core<br />

complex, and Psme1 (2.1 / 2.1) and Psme2 (2.7 / 3.1), which encode components <strong>of</strong> the regulatory<br />

11 S complex, were induced in both BALB/c and C57BL/6 BMM after IFN-γ treatment.<br />

96


Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

A B<br />

Fig. R.3.4: Ingenuity Pathway Analysis <strong>of</strong> IFN-γ effects in BALB/c and C57BL/6 BMM.<br />

EntrezGene ID lists <strong>of</strong> IFN-γ regulated genes were uploaded from Rosetta Resolver s<strong>of</strong>tware to Ingenuity Pathway Analysis (IPA, www.ingenuity.com). The two networks with the highest score from the separate<br />

analyses for BALB/c BMM and C57BL/6 BMM exhibited an overlap <strong>of</strong> 22 genes (nodes). Therefore, they covered similar functional aspects. Both networks were merged using the automatic merge-tool from IPA. The<br />

overlay data tool allows coloring <strong>of</strong> nodes corresponding to the fold change values from the comparisons “IFN-γ treated BALB/c BMM” vs. “non-treated medium control BALB/c BMM” (A) or “IFN-γ treated C57BL/6<br />

BMM” vs. “non-treated medium control C57BL/6 BMM” (B). Red indicates an increase while green marks a decrease <strong>of</strong> gene expression. The shade <strong>of</strong> the color correlates to the magnitude <strong>of</strong> change: more intense<br />

color shows a higher absolute fold change. Transcriptome expression data included only genes which passed the regulation criteria <strong>of</strong> ANOVA significance p* < 0.01 and the absolute fold change cut<strong>of</strong>f <strong>of</strong> 1.5.<br />

97


Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

The peptide transporters Tap1 and Tap2 translocate peptides from cytosolic proteasome into<br />

the endoplasmatic reticulum, where they bind to MHC-I molecules, while the peptide-MHCcomplexes<br />

are lateron transferred to the cell surface. Tap1 and Tap2 are described to be induced<br />

<strong>by</strong> IFN-γ (Brucet et al. 2004, e. g. human, Ma W et al. 1997, Schiffer et al. 2002). In this study,<br />

Tap1 (5.0 / 5.2) and Tap2 (3.3 / 3.0) were higher expressed after IFN-γ treatment in BMM <strong>of</strong> both<br />

strains BALB/c and C57BL/6 (Fig. R.3.4 A, B). Erap 1, an endoplasmic reticulum aminopeptidase<br />

that takes part in further processing and N-terminal trimming <strong>of</strong> the peptides in the ER (Chang et<br />

al. 2005, Jung et al. 2009), was additionally induced after IFN-γ treatment in BALB/c BMM (2.1)<br />

and C57BL/6 BMM (2.2) [data not shown]. Finally, several genes <strong>of</strong> MHC class I and class II<br />

molecules were induced after IFN-γ treatment in BMM <strong>of</strong> both strains BALB/c and C57BL/6 [data<br />

not shown].<br />

Comparison <strong>of</strong> differentially expressed gene lists (Fig. R.3.2 A), comparison <strong>of</strong> log 2 -ratio data<br />

(Fig. R.3.2 B), and Ingenuity Pathway Analyis (Fig. R.3.4) revealed highly similar gene expression<br />

signatures in BALB/c and C57BL/6 BMM after IFN-γ treatment even for genes which were only<br />

differentially expressed in one <strong>of</strong> the two comparisons. Therefore, further analysis <strong>of</strong> biological<br />

effects resulting from IFN-γ treatment was performed using the union list <strong>of</strong> IFN-γ effects in<br />

BALB/c and C57BL/6 BMM.<br />

First, the list <strong>of</strong> in total 531 IFN-γ dependent genes was subjected to a global functional<br />

analysis using Ingenuity Pathway Analysis (IPA, www.ingenuity.com), which compared functional<br />

categories within the data set with those <strong>of</strong> the complete array and assigned a p-value to rate on<br />

significant overrepresentation (Table R.3.3).<br />

Table R.3.3: Global functional analysis <strong>of</strong> IFN-γ dependent genes in BMM <strong>of</strong> at least one strain, BALB/c or C57BL/6, using Ingenuity<br />

Pathway Analysis (Ingenuity Systems, www.ingenuity.com).<br />

Diseases and Disorders<br />

category p-value a <strong>of</strong><br />

number<br />

genes<br />

Inflammatory<br />

Response<br />

Immunological<br />

Disease<br />

Inflammatory<br />

Disease<br />

Connective<br />

Tissue Disorders<br />

Skeletal and<br />

Muscular<br />

Disorders<br />

Infection<br />

Mechanism<br />

Cancer<br />

Respiratory<br />

Disease<br />

Antimicrobial<br />

Response<br />

Infectious<br />

Disease<br />

2.06E-06 -<br />

3.53E-40<br />

5.28E-06 -<br />

1.39E-23<br />

5.61E-06 -<br />

5.75E-23<br />

8.25E-07 -<br />

5.12E-21<br />

8.25E-07 -<br />

5.12E-21<br />

2.83E-06 -<br />

3.51E-19<br />

7.13E-06 -<br />

4.05E-17<br />

2.20E-06 -<br />

2.20E-16<br />

4.46E-08 -<br />

6.66E-15<br />

2.20E-06 -<br />

1.14E-14<br />

146<br />

150<br />

157<br />

Molecular and Cellular Functions<br />

category p-value a <strong>of</strong><br />

number<br />

genes<br />

Cellular Growth and<br />

Proliferation<br />

Cellular<br />

Development<br />

Cell-To-Cell<br />

Signaling and<br />

Interaction<br />

111 Cellular Movement<br />

150 Cell Death<br />

43<br />

141<br />

Cellular Function<br />

and Maintenance<br />

Antigen<br />

Presentation<br />

56 Cell Cycle<br />

29 Gene Expression<br />

87 Cell Signaling<br />

a Ten categories with the smallest p-values are cited.<br />

D. a. F. – Development and Function<br />

6.38E-06 -<br />

4.55E-25<br />

6.28E-06 -<br />

8.22E-25<br />

6.78E-06 -<br />

2.73E-23<br />

1.36E-09 -<br />

5.54E-21<br />

5.67E-06 -<br />

1.21E-18<br />

4.90E-06 -<br />

1.76E-18<br />

6.38E-06 -<br />

3.53E-16<br />

1.04E-06 -<br />

4.82E-10<br />

1.98E-07 -<br />

1.00E-09<br />

9.08E-07 -<br />

3.04E-09<br />

154<br />

117<br />

Physiological System<br />

Development and Function<br />

category p-value a <strong>of</strong><br />

number<br />

genes<br />

Hematological<br />

System D. a. F.<br />

Immune Cell<br />

Trafficking<br />

127 Hematopoiesis<br />

101<br />

150<br />

84<br />

56<br />

71<br />

69<br />

62<br />

Tissue<br />

Morphology<br />

Cell-mediated<br />

Immune<br />

Response<br />

Organismal<br />

Survival<br />

Tissue<br />

Development<br />

Humoral<br />

Immune<br />

Response<br />

Cardiovascular<br />

System D. a. F.<br />

Organismal<br />

Development<br />

6.38E-06 -<br />

4.55E-25<br />

4.90E-06 -<br />

6.53E-21<br />

2.06E-06 -<br />

2.65E-20<br />

9.94E-07 -<br />

1.36E-18<br />

4.01E-07 -<br />

2.27E-18<br />

2.50E-07 -<br />

3.64E-15<br />

6.82E-06 -<br />

4.43E-14<br />

9.94E-07 -<br />

2.03E-10<br />

4.05E-06 -<br />

2.39E-09<br />

3.05E-06 -<br />

9.33E-09<br />

138<br />

93<br />

82<br />

70<br />

66<br />

76<br />

71<br />

37<br />

39<br />

41<br />

98


Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

This analysis revealed expected categories within the IFN-γ dependently regulated genes in<br />

BMM like “Inflammatory Response”, “Antigen Presentation”, “Immune Cell Trafficking” and other<br />

immune response related functions. Furthermore, growth, proliferation, but also cell cyle and cell<br />

death associated genes were included in the set <strong>of</strong> differentially expressed genes.<br />

For a more detailed view, IPA’s canonical pathway analysis was applied (Table R.3.4). Here,<br />

the pathway “Interferon Signaling” appeared with the highest ratio value, i. e. the fraction <strong>of</strong><br />

genes from the pathway, which was also included in the list <strong>of</strong> IFN-γ dependent genes. This<br />

observation nicely proved the relevance <strong>of</strong> the gene expression signature which was recorded<br />

after stimulation <strong>of</strong> BMM with interferon. Further pathways from the group with the highest<br />

ratio values were among others “Antigen Presentation Pathway”, “Role <strong>of</strong> Pattern Recognition<br />

Receptors in Recognition <strong>of</strong> Bacteria and Viruses”, and “Activation <strong>of</strong> IRF <strong>by</strong> Cytosolic Pattern<br />

Recognition Receptors” which fits well into the expected <strong>characterization</strong> <strong>of</strong> the experimental<br />

system. Again, also cell death associated features appeared with the “Retinoic acid Mediated<br />

Apoptosis Signaling” pathway.<br />

Table R.3.4: Canonical pathway analysis <strong>of</strong> IFN-γ dependent genes in BMM <strong>of</strong> at least one strain, BALB/c or C57BL/6, using Ingenuity<br />

Pathway Analysis (Ingenuity Systems, www.ingenuity.com).<br />

Pathway ratio a p-value<br />

Interferon Signaling 11/30 = 0.367 8.34E-12<br />

Antigen Presentation Pathway 11/39 = 0.282 1.61E-09<br />

Role <strong>of</strong> Pattern Recognition Receptors in Recognition <strong>of</strong> Bacteria and Viruses 20/86 = 0.233 5.03E-15<br />

Complement System 8/36 = 0.222 4.09E-07<br />

Retinoic acid Mediated Apoptosis Signaling 9/44 = 0.205 2.42E-08<br />

Role <strong>of</strong> PKR in Interferon Induction and Antiviral Response 9/46 = 0.196 2.95E-07<br />

T Helper Cell Differentiation 8/41 = 0.195 7.42E-06<br />

Activation <strong>of</strong> IRF <strong>by</strong> Cytosolic Pattern Recognition Receptors 14/74 = 0.189 1.51E-11<br />

Allograft Rejection Signaling 8/45 = 0.178 2.34E-05<br />

TREM1 Signaling 12/69 = 0.174 3.84E-09<br />

a Ten pathways with the highest ratio values are cited.<br />

Manual data mining revealed different cytokines and cytokine receptors, which were<br />

differentially expressed in BMM upon treatment with IFN-γ. Here, the induction <strong>of</strong> several<br />

chemotactic chemokines (Ccl5, Ccl12, Ccl22, Cxcl9, Cxcl10, Cxcl11, Cxcl16) was noticeable. Among<br />

the receptors, Ccr5 induction fitted to the induction <strong>of</strong> one <strong>of</strong> its ligands, Ccl5. The induced<br />

receptor Ccrl2 is known to be induced after activation and during differentiation from monocytes<br />

to macrophages, but its function is not described yet. Contrarily, the receptor Cxcr4 was<br />

repressed after treatment <strong>of</strong> BMM with IFN-γ. Two interleukins were included in the<br />

differentially expressed cytokines: IL-15, a regulatory, anti-apoptotic cytokine with structural<br />

similarity to IL-2, and the subunit IL27-p28 <strong>of</strong> the heterodimeric IL-27, a regulatory cytokine<br />

produced <strong>by</strong> antigen presenting cells. Several interleukin receptors or receptor subunits were<br />

induced after IFN-γ treatment. The induction <strong>of</strong> the IL-15 receptor alpha-chain (Il15ra)<br />

corresponded to the already mentioned induction <strong>of</strong> IL-15. Furthermore, the strongly<br />

interdependent receptor subunits Il4ra, Il13ra1, Il21r, and Il2rg were induced. The alpha-chain <strong>of</strong><br />

IL-4 receptor (Il4ra) and the IL-21 receptor (Il21r) employ the IL-2 receptor gamma-chain (Il2rg) as<br />

their second subunit. Similarly, the primary IL-13 binding subunit <strong>of</strong> the IL-13 receptor (Il13ra1)<br />

builds a receptor complex together with the alpha-chain <strong>of</strong> IL-4 receptor (Il4ra). Also Il12rb1,<br />

coding for the low-affinity binding subunit <strong>of</strong> the IL-12 receptor, was induced. The induction <strong>of</strong><br />

Il10ra, high-affinity binding subunit <strong>of</strong> the IL-10 receptor, and <strong>of</strong> Il18bp, cytokine IL-18 binding<br />

protein, is an example for inhibitory processes in BMM after IFN-γ treatment. Repression <strong>of</strong><br />

interleukin receptor subunits was also found after IFN-γ treatment: Il1rl2, receptor for interleukin<br />

1 family member 9, and Il6st alias gp130, signal transducer for several cytokines, e. g. IL-6, were<br />

99


Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

found to be repressed. Finally, differential expression was observed for members <strong>of</strong> the TNF<br />

ligand and receptor superfamilies. The central inflammation mediator Tnf, the apoptosis inducer<br />

TRAIL (Tnfsf10), and the T cell survival modulator Tnfsf18 were induced in parallel with the<br />

receptor Tnfrsf14 (Table R.3.5).<br />

Table R.3.5: IFN-γ influence on gene expression <strong>of</strong> cytokines and cytokine receptors in BMM <strong>of</strong> BALB/c and C57BL/6 mice.<br />

Rosetta Resolver Annotation fold change a strain<br />

difference<br />

gene<br />

Entrez<br />

description<br />

alias<br />

name<br />

Gene ID<br />

BALB/c C57BL/6 control IFN-γ<br />

Ccl5 chemokine (C-C motif) ligand 5<br />

SISd, Scya5, RANTES,<br />

TCP228<br />

20304 3.7 5.0<br />

Ccl12 chemokine (C-C motif) ligand 12 MCP-5, Scya12 20293 3.4 2.1<br />

Ccl22 chemokine (C-C motif) ligand 22<br />

MDC, DCBCK, ABCD-1,<br />

Scya22<br />

20299 2.9 3.6<br />

Ccr5 chemokine (C-C motif) receptor 5 AM4-7, CD195, Cmkbr5 12774 1.7 1.4<br />

Ccrl2 chemokine (C-C motif) receptor-like 2<br />

E01, CCR11, L-CCR,<br />

Cmkbr1l2<br />

54199 2.0 1.7<br />

Cxcl9 chemokine (C-X-C motif) ligand 9 CMK, Mig, Scyb9, crg-10 17329 45.0 56.1<br />

Cxcl10 chemokine (C-X-C motif) ligand 10<br />

C7, IP10, CRG-2, INP10,<br />

Ifi10, mob-1, Scyb10, 15945 47.8 21.5<br />

gIP-10<br />

Cxcl11 chemokine (C-X-C motif) ligand 11<br />

IP9, H174, ITAC, CXC11,<br />

SCYB9B, Scyb11, betaR1<br />

56066 8.3 1.1 x<br />

Cxcl16 chemokine (C-X-C motif) ligand 16 SR-PSOX, Zmynd15 66102 2.6 2.9<br />

Cxcr4 chemokine (C-X-C motif) receptor 4<br />

CD184, LESTR, Sdf1r,<br />

Cmkar4, PB-CKR, 12767 -1.8 -2.2 x x<br />

PBSF/SDF-1<br />

Il1rl2 interleukin 1 receptor-like 2 IL-1Rrp2 107527 -1.7 -1.8<br />

Il2rg interleukin 2 receptor, gamma chain CD132, gamma(c) 16186 1.9 1.7<br />

Il4ra interleukin 4 receptor, alpha Il4r, CD124 16190 2.0 1.6 x x<br />

Il6st interleukin 6 signal transducer CD130, gp130 16195 -1.8 -1.8<br />

Il10ra interleukin 10 receptor, alpha Il10r, CDw210, mIL-10R 16154 1.6 1.7<br />

Il12rb1 interleukin 12 receptor, beta 1 CD212, IL-12R[b] 16161 1.8 3.9<br />

Il13ra1 interleukin 13 receptor, alpha 1 NR4, Il13ra, CD213a1 16164 2.1 1.7 x<br />

Il15 interleukin 15 16168 2.8 2.0 x<br />

Il15ra interleukin 15 receptor, alpha chain 16169 4.3 4.0<br />

Il18bp interleukin 18 binding protein MC54L, Igifbp, IL-18BP 16068 5.5 6.2<br />

Il21r interleukin 21 receptor NILR 60504 3.4 2.8 x x<br />

Il27 interleukin 27 p28, Il30, IL-27p28 246779 2.3 2.4<br />

Tnf tumor necrosis factor<br />

DIF, Tnfa, TNFSF2,<br />

Tnfsf1a, TNF-alpha<br />

21926 2.4 2.7<br />

Tnfsf10<br />

tumor necrosis factor (ligand) superfamily,<br />

member 10<br />

TL2, Ly81, Trail, APO-2L 22035 10.6 5.2<br />

Tnfsf18<br />

tumor necrosis factor (ligand) superfamily,<br />

member 18<br />

Gitrl 240873 3.3 1.1 x<br />

Tnfrsf14<br />

tumor necrosis factor receptor superfamily, Atar, HveA, Hvem,<br />

member 14 (herpesvirus entry mediator) Tnfrs14<br />

230979 2.6 2.5<br />

a Fold change values were calculated from expression intensities <strong>of</strong> IFN-γ treated BMM in comparison to control BMM from the mean<br />

<strong>of</strong> three biological replicates. Differential expression in statistical testing with p* < 0.01 and a minimal absolute fold change <strong>of</strong> 1.5 is<br />

indicated in bold.<br />

Furthermore, GTPases / GTP binding proteins <strong>of</strong> different families, which are known to be<br />

induced <strong>by</strong> interferon and play a role in antimicrobial defense, were induced, and some even<br />

belonged to the genes with the strongest induction within the data set. Four families are<br />

described in literature (MacMicking 2004), p47 GTPases, p65 guanylate-binding proteins (GBP),<br />

Mx proteins and very large inducible GTPases (VLIG), and <strong>of</strong> each family examples were included<br />

in the data set <strong>of</strong> induced genes in BMM after IFN-γ treatment. The p47 family was present with<br />

the members Igtp, Iigp1, Irgm1, and Tgtp. All guanylate binding proteins <strong>of</strong> the p65 family were<br />

induced (Gbp1, Gbp2, Gbp3, Gbp4, Gbp5, and Gbp6). Both types <strong>of</strong> Mx genes, coding for the<br />

nuclear (Mx1) and the cytosolic (Mx2) form, exhibited increased expression in BMM after IFN-γ<br />

treatment. Finally, very large interferon inducible GTPase 1, Gvin1, was included in the list <strong>of</strong><br />

induced genes as member <strong>of</strong> the last family (Table R.3.6).<br />

100


Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

Table R.3.6: IFN-γ influence on gene expression <strong>of</strong> GTPase family members in BMM <strong>of</strong> BALB/c and C57BL/6 mice.<br />

Rosetta Resolver Annotation fold change a strain<br />

difference<br />

gene<br />

Entrez<br />

description<br />

alias<br />

name<br />

Gene ID<br />

BALB/c C57BL/6 control IFN-γ<br />

Igtp interferon gamma induced GTPase Irgm3 16145 17.4 13.5<br />

Iigp1 interferon inducible GTPase 1 Iigp, Irga6 60440 100.0 100.0<br />

Irgm1 immunity-related GTPase family M member 1<br />

Ifi1, Irgm, Iigp3, Iipg3,<br />

LRG-47<br />

15944 6.9 11.2<br />

Tgtp T-cell specific GTPase Tgtp, Gtp2, Mg21, Irgb6 21822 49.0 100.0<br />

Gbp1 guanylate binding protein 1 Mpa1, Gbp-1, Mag-1 14468 43.9 6.0 x x<br />

Gbp2 guanylate binding protein 2 14469 17.2 31.2 x<br />

Gbp3 guanylate binding protein 3 Gbp4 55932 21.7 15.6<br />

Gbp4 guanylate binding protein 4 Mpa2, Mag-2, Mpa-2 17472 60.6 94.3<br />

Gbp5 guanylate binding protein 5 Gbp5a 229898 50.7 66.8<br />

Gbp6 guanylate binding protein 6 Gbp7 229900 9.3 7.2<br />

Mx1 myxovirus (influenza virus) resistance 1 Mx, Mx-1 17857 10.8 6.9 x<br />

Mx2 myxovirus (influenza virus) resistance 2 Mx-2 17858 6.1 3.5 x<br />

Gvin1 GTPase, very large interferon inducible 1 VLIG, Iigs1, VLIG-1 74558 4.6 6.7<br />

a Fold change values were calculated from expression intensities <strong>of</strong> IFN-γ treated BMM in comparison to control BMM from the mean<br />

<strong>of</strong> three biological replicates. Differential expression in statistical testing with p* < 0.01 and a minimal absolute fold change <strong>of</strong> 1.5 is<br />

indicated in bold.<br />

BMM after IFN-γ treatment induced the expression <strong>of</strong> complement system components. Of<br />

the classical activation pathway, the C1 subcomponents C1qa, C1qb, C1qc, C1r, and C1rb were<br />

induced, and <strong>of</strong> the alternative pathway factor B (Cfb). Induced component C3 takes part in all<br />

activation pathways and especially in the amplification loop <strong>of</strong> complement activation<br />

(Table R.3.7).<br />

Table R.3.7: IFN-γ influence on gene expression <strong>of</strong> complement components in BMM <strong>of</strong> BALB/c and C57BL/6 mice.<br />

Rosetta Resolver Annotation fold change a strain<br />

difference<br />

gene<br />

Entrez<br />

description<br />

alias<br />

name<br />

Gene ID<br />

BALB/c C57BL/6 control IFN-γ<br />

C1qa<br />

complement component 1, q subcomponent, alpha<br />

polypeptide<br />

C1q 12259 1.3 2.3<br />

C1qb<br />

complement component 1, q subcomponent, beta<br />

polypeptide<br />

12260 2.2 2.8 x x<br />

C1qc complement component 1, q subcomponent, C chain C1qg, Ciqc 12262 1.6 2.8 x<br />

C1r complement component 1, r subcomponent C1ra, C1rb 50909 4.3 4.7<br />

C1rb complement component 1, r subcomponent, b 270510 5.5 6.3<br />

C3 complement component 3 ASP, Plp 12266 4.9 5.6<br />

Cfb complement factor B B, Bf, Fb, H2-Bf 14962 14.1 19.5<br />

a Fold change values were calculated from expression intensities <strong>of</strong> IFN-γ treated BMM in comparison to control BMM from the mean<br />

<strong>of</strong> three biological replicates. Differential expression in statistical testing with p* < 0.01 and a minimal absolute fold change <strong>of</strong> 1.5 is<br />

indicated in bold.<br />

Other immune function related gene expression changes were conspicuous in BMM after<br />

IFN-γ treatment. Inducible nitric oxide synthase 2 (Nos2) was increased as expected in treated<br />

BMM. The enzyme produces NO, which acts as antimicrobial effector and as signal mediator<br />

during immune defense processes. Another induced enzyme gene was kynureninase (Lkynurenine<br />

hydrolase, Kynu) whose product is involved in the degradation <strong>of</strong> kynurenines, toxic<br />

intermediates with signaling properties. Their production is amplified in inflammation settings <strong>by</strong><br />

the induction <strong>of</strong> indoleamine 2,3-dioxygenase (IDO), which catalyzes the first step <strong>of</strong> tryptophan<br />

degradation in the kyurenine pathway. Interestingly, the Ido gene was not differentially<br />

expressed in BMM after IFN-γ treatment, expression was even absent in the majority <strong>of</strong> samples.<br />

The tryptophanyl-tRNA synthetase (Wars) was induced in IFN-γ treated BMM. Prostaglandin-<br />

101


Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

endoperoxide synthase 2 (Ptgs2), coding for the enzyme responsible for prostaglandin E 2<br />

biosynthesis, was induced as well as different 2'-5'-oligoadenylate synthetase genes (Oas2, Oas3,<br />

Oasl1, Oasl2). They are part <strong>of</strong> the innate immune response. Serum amyloid proteins are part <strong>of</strong><br />

the acute phase response. In humans, Saa genes are known to be expressed not only in the liver,<br />

but also in activated monocytes/macrophages (Urieli-Shoval et al. 1994). In this study, serum<br />

amyloid A 3 (Saa3) was induced in murine BMM after IFN-γ treatment.<br />

BMM induced plasma membrane receptors after treatment with IFN-γ. Among others, the<br />

induction <strong>of</strong> IgG Fc receptor with high affinity I and with low affinity IV (Fcgr1 and Fcgr4), <strong>of</strong> tolllike<br />

receptors 2, 9, and 12 (Tlr2, Tlr9, Tlr12), and <strong>of</strong> co-stimulatory receptors Cd40 and Cd86 was<br />

observed. Additionally, CD274 (B7-H1, PD-L1), ligand for the immunoinhibitory receptor PD-1 on<br />

activated B and T cells, and the second PD-1 receptor, programmed cell death 1 ligand 2<br />

(Pdcd1lg2; PD-L2) exhibited induction (Table R.3.8).<br />

Table R.3.8: IFN-γ influence on expression <strong>of</strong> immune function related genes in BMM <strong>of</strong> BALB/c and C57BL/6 mice.<br />

Rosetta Resolver Annotation fold change a strain<br />

difference<br />

gene<br />

Entrez<br />

description<br />

alias<br />

name<br />

Gene ID<br />

BALB/c C57BL/6 control IFN-γ<br />

Nos2 nitric oxide synthase 2, inducible iNOS,Nos2a 18126 18.6 25.5<br />

Kynu kynureninase (L-kynurenine hydrolase) 70789 4.0 5.2<br />

Wars tryptophanyl-tRNA synthetase WRS 22375 3.7 4.4<br />

Ptgs2 prostaglandin-endoperoxide synthase 2<br />

COX2, PHS-2, Pghs2,<br />

TIS10, PGHS-2<br />

19225 16.2 15.4<br />

Oas2 2'-5' oligoadenylate synthetase 2 Oasl11 246728 2.3 2.4 x<br />

Oas3 2'-5' oligoadenylate synthetase 3 Oasl10 246727 3.0 4.4<br />

Oasl1 2'-5' oligoadenylate synthetase-like 1 oasl9 231655 4.8 2.4 x x<br />

Oasl2 2'-5' oligoadenylate synthetase-like 2<br />

Oasl, M1204, Mmu-<br />

OASL<br />

23962 2.6 4.2<br />

Saa3 serum amyloid A 3 l7R3, Saa-3 20210 3.8 5.7<br />

Fcgr1 Fc receptor, IgG, high affinity I<br />

CD64, IGGHAFC,<br />

FcgammaRI<br />

14129 4.6 4.9 x<br />

Fcgr4 Fc receptor, IgG, low affinity IV<br />

Fcrl3, CD16-2, FcgRIV,<br />

Fcgr3a, FcgammaRIV<br />

246256 6.0 3.9<br />

Tlr2 toll-like receptor 2 Ly105 24088 2.0 1.5<br />

Tlr9 toll-like receptor 9 81897 2.8 2.3<br />

Tlr12 toll-like receptor 12 Gm1365 384059 1.8 2.8<br />

Cd40 CD40 antigen<br />

IGM, p50, Bp50, GP39,<br />

IMD3, TRAP, HIGM1, T- 21939 5.1 2.6 x<br />

BAM, Tnfrsf5<br />

Cd86 CD86 antigen<br />

B7, B70, MB7, B7-2,<br />

B7.2, CLS1, Ly58, ETC-1,<br />

Ly-58, MB7-2, Cd28l2,<br />

12524 6.5 5.4<br />

TS/A-2<br />

Cd274 CD274 antigen<br />

B7-H1, PD-L1, Pdcd1l1,<br />

Pdcd1lg1<br />

60533 4.0 3.0<br />

Pdcd1lg2 programmed cell death 1 ligand 2 Btdc, B7-DC, PD-L2 58205 5.5 8.6<br />

a Fold change values were calculated from expression intensities <strong>of</strong> IFN-γ treated BMM in comparison to control BMM from the mean<br />

<strong>of</strong> three biological replicates. Differential expression in statistical testing with p* < 0.01 and a minimal absolute fold change <strong>of</strong> 1.5 is<br />

indicated in bold.<br />

Finally, the expression <strong>of</strong> mitochondrial superoxide dismutase 2 (Sod2) and glutaredoxin<br />

(Glrx), which are involved in the anti-oxidant defense, was induced. Contrarily, glutathione<br />

S-transferase mu 1 (Gstm1), which is associated to detoxification processes, was repressed.<br />

Furthermore, the induction <strong>of</strong> lysosomal enzymes cathepsin C and H (Ctsc, Ctsh) was observed.<br />

BMM induced cell adhesion molecules, integrins, and protocadherin (Icam1, Itgal, Itgb7, Pcdh7,<br />

Vcam1) after IFN-γ treatment, whereas they repressed the beta-integrin subunit Itgb3<br />

(Table R.3.9).<br />

102


Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

Table R.3.9: IFN-γ influence on gene expression <strong>of</strong> anti-oxidant, detoxification, and lysosomal enzymes and <strong>of</strong> adhesion molecules in<br />

BMM <strong>of</strong> BALB/c and C57BL/6 mice.<br />

Rosetta Resolver Annotation fold change a strain<br />

difference<br />

gene<br />

Entrez<br />

description<br />

alias<br />

name<br />

Gene ID<br />

BALB/c C57BL/6 control IFN-γ<br />

Sod2 superoxide dismutase 2, mitochondrial MnSOD 20656 2.1 2.2<br />

Glrx glutaredoxin Grx1, Glrx1, Ttase 93692 2.0 2.2<br />

Gstm1 glutathione S-transferase, mu 1 Gstb1, Gstb-1 14862 -1.9 -1.9<br />

Ctsc cathepsin C DPP1, DPPI 13032 4.0 4.5 x<br />

Ctsh cathepsin H Ctsh 13036 2.2 1.6 x x<br />

Icam1 intercellular adhesion molecule 1 CD54, Ly-47, Icam-1, MALA-2 15894 2.3 2.2<br />

Itgal integrin alpha L Cd11a, LFA-1, Ly-15, Ly-21 16408 2.8 2.0 x<br />

Itgb3 integrin beta 3 CD61, GP3A, INGRB3 16416 -2.3 -2.9<br />

Itgb7 integrin beta 7 Ly69 16421 2.5 2.1<br />

Pcdh7 protocadherin 7 54216 1.7 1.7 x<br />

Vcam1 vascular cell adhesion molecule 1 CD106, Vcam-1 22329 3.1 1.7<br />

a Fold change values were calculated from expression intensities <strong>of</strong> IFN-γ treated BMM in comparison to control BMM from the mean<br />

<strong>of</strong> three biological replicates. Differential expression in statistical testing with p* < 0.01 and a minimal absolute fold change <strong>of</strong> 1.5 is<br />

indicated in bold.<br />

Biological context <strong>of</strong> gene expression differences between BALB/c BMM and C57BL/6 BMM<br />

Comparison <strong>of</strong> differentially expressed gene lists (Fig. R.3.3 A) and comparison <strong>of</strong> log 2 -ratio<br />

data (Fig. R.3.3 B) revealed highly similar strain differences between BALB/c and C57BL/6 BMM at<br />

both treatment levels, non-treated medium control and after IFN-γ treatment, even for genes<br />

which were differentially expressed only in one <strong>of</strong> the two comparisons. Therefore, the biological<br />

context <strong>of</strong> strain differences was further analyzed using the union set <strong>of</strong> differences at control<br />

level and after IFN-γ treatment.<br />

First, the list <strong>of</strong> in total 324 genes, which exhibited differences between the strains in at least<br />

one <strong>of</strong> the two treatment conditions, was subjected to a global functional analysis using<br />

Ingenuity Pathway Analysis (IPA, www.ingenuity.com), which compared functional categories<br />

within the data set with those <strong>of</strong> the complete array and assigned a p-value to rate on significant<br />

overrepresentation (Table R.3.10).<br />

This analysis revealed on the one hand categories with immune response related functions<br />

like “Inflammatory Response”, “Antigen Presentation”, and “Immune Cell Trafficking”, which was<br />

not surprising because the gene expression repertoire <strong>of</strong> BMM independent from the strain <strong>of</strong><br />

which the cells were derived is focused at the immune response as the proprietary function <strong>of</strong><br />

the BMM. Metabolic functions and the category “Cell Death” additionally appeared in the<br />

analysis <strong>of</strong> strain differences.<br />

Also for the strain difference aspect <strong>of</strong> this study, IPA’s canonical pathway analysis was<br />

applied for a more detailed view (Table R.3.11).<br />

Here, the pathway “Complement System” appeared with the highest ratio value. Four genes<br />

differentially expressed between the strains were included in this pathway. Noticeably, the ratio<br />

values for this and further pathways were lower than in the analysis <strong>of</strong> IFN-γ effects on BMM.<br />

This might reflect the smaller number <strong>of</strong> genes exhibiting differences between the strains in<br />

comparison to the number <strong>of</strong> genes which are influenced <strong>by</strong> IFN-γ, but also the possibility that<br />

the strain differences are more distributed among the pathways and functions.<br />

103


Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

Table R.3.10: Global functional analysis <strong>of</strong> genes exhibiting strain differences in BMM <strong>of</strong> at least one treatment group, medium control<br />

or after IFN-γ treatment, using Ingenuity Pathway Analysis (Ingenuity Systems, www.ingenuity.com).<br />

Diseases and Disorders<br />

category p-value a <strong>of</strong><br />

number<br />

genes<br />

Inflammatory<br />

Response<br />

Cancer<br />

Antimicrobial<br />

Response<br />

Inflammatory<br />

Disease<br />

Genetic Disorder<br />

Connective<br />

Tissue Disorders<br />

Skeletal and<br />

Muscular<br />

Disorders<br />

Gastrointestinal<br />

Disease<br />

Metabolic<br />

Disease<br />

Immunological<br />

Disease<br />

1.51E-02 -<br />

2.19E-14<br />

1.51E-02 -<br />

6.88E-14<br />

5.60E-05 -<br />

6.53E-11<br />

1.51E-02 -<br />

1.09E-10<br />

1.51E-02 -<br />

3.99E-08<br />

1.51E-02 -<br />

4.61E-08<br />

1.51E-02 -<br />

4.61E-08<br />

1.51E-02 -<br />

6.60E-07<br />

1.51E-02 -<br />

8.95E-07<br />

1.51E-02 -<br />

4.22E-06<br />

69<br />

Molecular and Cellular Functions<br />

category p-value a <strong>of</strong><br />

number<br />

genes<br />

Cellular Growth and<br />

Proliferation<br />

95 Lipid Metabolism<br />

18<br />

Small Molecule<br />

Biochemistry<br />

88 Cell Death<br />

146 Cellular Movement<br />

58<br />

78<br />

Cell-To-Cell<br />

Signaling and<br />

Interaction<br />

Antigen<br />

Presentation<br />

70 Cell Morphology<br />

a Ten categories with the smallest p-values are cited.<br />

D. a. F. – Development and Function<br />

75<br />

66<br />

Cellular<br />

Development<br />

Carbohydrate<br />

Metabolism<br />

1.36E-02 -<br />

2.43E-08<br />

1.51E-02 -<br />

4.00E-08<br />

1.51E-02 -<br />

4.00E-08<br />

8.68E-03 -<br />

5.49E-07<br />

1.51E-02 -<br />

5.53E-07<br />

1.51E-02 -<br />

1.14E-06<br />

1.19E-02 -<br />

1.08E-05<br />

1.51E-02 -<br />

1.86E-05<br />

1.51E-02 -<br />

1.86E-05<br />

1.51E-02 -<br />

3.81E-05<br />

92<br />

28<br />

42<br />

81<br />

47<br />

64<br />

24<br />

38<br />

44<br />

Physiological System<br />

Development and Function<br />

category p-value a <strong>of</strong><br />

number<br />

genes<br />

Hematological<br />

System D. a. F.<br />

Immune Cell<br />

Trafficking<br />

Cardiovascular<br />

System D. a. F.<br />

Organismal<br />

Development<br />

Humoral<br />

Immune<br />

Response<br />

Organismal<br />

Survival<br />

Cell-mediated<br />

Immune<br />

Response<br />

Tissue<br />

Development<br />

Tissue<br />

Morphology<br />

20 Hematopoiesis<br />

1.51E-02 -<br />

7.63E-07<br />

1.37E-02 -<br />

7.63E-07<br />

1.51E-02 -<br />

3.45E-05<br />

9.01E-03 -<br />

3.45E-05<br />

1.51E-02 -<br />

5.60E-05<br />

1.45E-02 -<br />

6.36E-05<br />

1.51E-02 -<br />

7.36E-05<br />

1.51E-02 -<br />

1.43E-04<br />

1.37E-02 -<br />

2.29E-04<br />

1.51E-02 -<br />

7.48E-04<br />

65<br />

38<br />

30<br />

18<br />

14<br />

19<br />

13<br />

29<br />

35<br />

35<br />

Table R.3.11: Canonical pathway analysis <strong>of</strong> genes exhibiting strain differences in BMM <strong>of</strong> at least one treatment group, medium<br />

control or after IFN-γ treatment, using Ingenuity Pathway Analysis (Ingenuity Systems, www.ingenuity.com).<br />

Pathway ratio a p-value<br />

Complement System 4/36 = 0,111 1,04E-03<br />

Crosstalk between Dendritic Cells and Natural Killer Cells 8/98 = 0,082 1,45E-05<br />

Role <strong>of</strong> Pattern Recognition Receptors in Recognition <strong>of</strong> Bacteria and Viruses 7/86 = 0,081 1,08E-04<br />

Antigen Presentation Pathway 3/39 = 0,077 4,87E-03<br />

Allograft Rejection Signaling 3/45 = 0,067 6,18E-03<br />

Autoimmune Thyroid Diseases Signaling 3/47 = 0,064 7,69E-03<br />

Caveolar-mediated Endocytosis Signaling 5/83 = 0,060 4,44E-03<br />

Role <strong>of</strong> RIG1-like receptors in Antiviral Innate Immunity 3/52 = 0,058 1,34E-02<br />

IL-10 Signaling 4/70 = 0,057 1,02E-02<br />

Toll-like Receptor Signaling 3/54 = 0,056 3,06E-02<br />

a Ten pathways with the highest ratio values which additionally are over-represented with p < 0.05 are cited.<br />

Manual data mining revealed different cytokines and cytokine receptors, which were<br />

differentially expressed between BMM <strong>of</strong> the two strain at control level, upon treatment with<br />

IFN-γ, or in both conditions (Table R.3.12).<br />

Certain chemotactic chemokines were higher expressed in BALB/c BMM (Ccl24, Cxcl11,<br />

Cxcl14), but also for the receptor Ccr2, which binds Ccl2, Ccl7, and Ccl13 chemokines, and Cxcr4,<br />

whose ligand is Cxcl12, higher expression was observed in BALB/c BMM. Additionally, IL-15, a<br />

104


Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

regulatory, anti-apoptotic cytokine with structural similarity to IL-2, and the immune modulators<br />

Tnfsf8 and Tnfsf18, members <strong>of</strong> the TNF ligand superfamily, were higher expressed in BALB/c<br />

BMM.<br />

Strain differences were observed for the interleukin receptors Il4ra (alpha-chain <strong>of</strong> IL-4<br />

receptor), Il13ra1 (primary IL-13 binding subunit <strong>of</strong> the IL-13 receptor), and Il21r (IL-21 receptor).<br />

Il4ra and Il13ra1 are linked, because Il13ra1 builds a receptor complex together with Il4ra. All<br />

three interleukin receptor chains were higher expressed in C57BL/6 BMM (Table R.3.12).<br />

Table R.3.12: Strain differences in gene expression <strong>of</strong> cytokines and cytokine receptors between BMM <strong>of</strong> BALB/c and C57BL/6 mice.<br />

Rosetta Resolver Annotation fold change a IFN-γ effect<br />

gene<br />

Entrez<br />

description<br />

alias<br />

name<br />

Gene ID<br />

control IFN-γ BALB/c C57BL/6<br />

Ccl24 chemokine (C-C motif) ligand 24 CKb-6, MPIF-2, Scya24 56221 -2.7 -2.2<br />

Ccr2 chemokine (C-C motif) receptor 2<br />

Ckr2, Ccr2a, Ccr2b,<br />

Ckr2a, Ckr2b, mJe-r, 12772 -2.8 -4.0<br />

Cmkbr2, Cc-ckr-2<br />

Cxcl11 chemokine (C-X-C motif) ligand 11<br />

IP9, H174, ITAC, b-R1,<br />

CXC11, I-TAC, SCYB9B, 56066 -1.0 -7.6 x<br />

Scyb11, betaR1<br />

Cxcl14 chemokine (C-X-C motif) ligand 14<br />

KS1, Kec, BMAC, BRAK,<br />

NJAC, MIP-2g, Scyb14, 57266 -4.0 -6.1<br />

bolekine, MIP2gamma<br />

Cxcr4 chemokine (C-X-C motif) receptor 4<br />

CD184, LESTR, Sdf1r,<br />

Cmkar4, PB-CKR, 12767 -2.1 -2.5 x<br />

PBSF/SDF-1<br />

Il4ra interleukin 4 receptor, alpha Il4r, CD124 16190 2.2 1.7 x x<br />

Il13ra1 interleukin 13 receptor, alpha 1 NR4, Il13ra, CD213a1 16164 1.7 1.4 x x<br />

Il15 interleukin 15 16168 -1.5 -2.0 x<br />

Il21r interleukin 21 receptor NILR 60504 3.3 2.8 x<br />

Tnfsf18<br />

tumor necrosis factor (ligand) superfamily,<br />

member 18<br />

Gitrl 240873 1.0 -3.0 x<br />

Tnfsf8<br />

tumor necrosis factor (ligand) superfamily,<br />

member 8<br />

CD153, Cd30l, CD30LG 21949 -1.8 -5.2<br />

a Fold change values were calculated for the comparison <strong>of</strong> C57BL/6 BMM with the baseline <strong>of</strong> BALB/c BMM <strong>of</strong> mean values from<br />

three biological replicates per group. Positive values indicate higher expression in C57BL/6 BMM than in BALB/c BMM, whereas<br />

negative values indicate higher expression in BALB/c BMM than in C57BL/6 BMM. Differential expression between BMM <strong>of</strong> the two<br />

strains at the same treatment level (IFN-γ treated or medium control) using statistical testing in combination with an absolute fold<br />

change cut<strong>of</strong>f <strong>of</strong> 1.5 is indicated in bold.<br />

Furthermore, four GTPases / GTP binding proteins <strong>of</strong> the p65 and the Mx families, which were<br />

already observed to be induced <strong>by</strong> IFN-γ within the data set <strong>of</strong> BALB/c and C57BL/6 BMM,<br />

exhibited strain differences. Gbp1, Mx1, and Mx2 were higher expressed in BALB/c BMM,<br />

whereas Gbp2 was higher expressed in C57BL/6 BMM.<br />

Complement system components C1qb and C1qc, the receptor for complement factor<br />

fragment C5a, C5ar1, but also the complement factor C1 inhibitor, Serping1, were higher<br />

expressed in C57BL/6 BMM (Table R.3.13).<br />

Other immune function related gene expression differences between BMM <strong>of</strong> the two strains<br />

were conspicuous, because the genes were clearly affected <strong>by</strong> IFN-γ treatment. First, MHC genes<br />

<strong>of</strong> class I (H28, H2-Q6, H2-Q8, H2-T24) and class II (H2-Ea) were higher expressed in BALB/c BMM.<br />

But also H2-Ab1 (class II), H2-K1, and H2-M2 (class I) exhibited strain differences, however, their<br />

expression was higher in C57BL/6 BMM. Second, a group <strong>of</strong> interferon-inducible genes possessed<br />

differences in their expression between BMM <strong>of</strong> both strains: Ifi27l2a, Ifi44, Ifi202, Ifi203, Ifih1,<br />

Ifit3, Ifitm3, Irf7, and Isg20. All <strong>of</strong> them were higher expressed in BALB/c BMM than in C57BL/6<br />

BMM (Table R.3.14).<br />

105


Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

Table R.3.13: Strain differences in gene expression <strong>of</strong> GTPase family members and complement system components between BMM <strong>of</strong><br />

BALB/c and C57BL/6 mice.<br />

Rosetta Resolver Annotation fold change a IFN-γ effect<br />

gene<br />

Entrez<br />

description<br />

alias<br />

name<br />

Gene ID<br />

control IFN-γ BALB/c C57BL/6<br />

Gbp1 guanylate binding protein 1 Mpa1, Gbp-1, Mag-1 14468 -4.1 -30.0 x x<br />

Gbp2 guanylate binding protein 2 14469 1.3 2.3 x x<br />

Mx1 myxovirus (influenza virus) resistance 1 Mx, Mx-1 17857 -2.0 -3.1 x x<br />

Mx2 myxovirus (influenza virus) resistance 2 Mx-2 17858 -3.0 -5.3 x x<br />

C1qb<br />

complement component 1, q subcomponent,<br />

beta polypeptide<br />

12260 6.2 7.8 x<br />

C1qc<br />

complement component 1, q subcomponent,<br />

C chain<br />

C1qg, Ciqc 12262 2.0 3.5 x<br />

C5ar1 complement component 5a receptor 1 C5r1, Cd88, D7Msu1 12273 2.9 2.1<br />

Serping1<br />

serine (or cysteine) peptidase inhibitor, clade G,<br />

C1nh, C1INH<br />

member 1<br />

12258 1.1 2.1 x<br />

a Fold change values were calculated for the comparison <strong>of</strong> C57BL/6 BMM with the baseline <strong>of</strong> BALB/c BMM <strong>of</strong> mean values from<br />

three biological replicates per group. Positive values indicate higher expression in C57BL/6 BMM than in BALB/c BMM, whereas<br />

negative values indicate higher expression in BALB/c BMM than in C57BL/6 BMM. Differential expression between BMM <strong>of</strong> the two<br />

strains at the same treatment level (IFN-γ treated or medium control) using statistical testing in combination with an absolute fold<br />

change cut<strong>of</strong>f <strong>of</strong> 1.5 is indicated in bold.<br />

Table R.3.14: Strain differences in gene expression <strong>of</strong> MHC molecules and interferon-inducible factors between BMM <strong>of</strong> BALB/c and<br />

C57BL/6 mice.<br />

Rosetta Resolver Annotation fold change a IFN-γ effect<br />

gene<br />

Entrez<br />

description<br />

alias<br />

name<br />

Gene ID<br />

control IFN-γ BALB/c C57BL/6<br />

H28 histocompatibility 28 H-28, H28-1, NS1178 15061 -21.1 -100.0 x<br />

H2-Ea histocompatibility 2, class II antigen E alpha<br />

Ia3, Ia-3, H-2Ea,<br />

AI323765, E-alpha-f<br />

14968 -64.2 -100.0 x<br />

H2-Q6 histocompatibility 2, Q region locus 6 Qa6, Qa-6, H-2Q6 110557 -2.8 -1.9 x x<br />

H2-Q8 histocompatibility 2, Q region locus 8<br />

Qa8, Qa-2, H-2Q8,<br />

Ms10t, MMS10-T<br />

15019 -4.8 -3.2 x<br />

H2-T24 histocompatibility 2, T region locus 24 H-2T24 15042 -4.6 -2.5 x x<br />

H2-Ab1 histocompatibility 2, class II antigen A, beta 1<br />

IAb, Ia2, Ia-2, Abeta,<br />

H-2Ab, H2-Ab, Rmcs1, 14961 2.1 2.0<br />

I-Abeta<br />

H2-K1 histocompatibility 2, K1, K region H-2K, H2-K, H-2K(d) 14972 1.5 1.6<br />

H2-M2 histocompatibility 2, M region locus 2 H-2M2, Thy19.4 14990 6.2 4.4 x x<br />

Ifi27l2a interferon, alpha-inducible protein 27 like 2A Ifi27, Isg12, Isg12(b1) 76933 -5.4 -4.0<br />

Ifi44 interferon-induced protein 44 p44, MTAP44 99899 -36.1 -5.4 x x<br />

Ifi202 interferon activated gene 202 15949 -20.1 -78.6 x<br />

Ifi203 interferon activated gene 203 15950 -3.9 -2.0 x x<br />

Ifih1 interferon induced with helicase C domain 1<br />

Hlcd, MDA5, MDA-5,<br />

Helicard<br />

71586 -1.5 -1.9 x x<br />

Ifit3<br />

interferon-induced protein with<br />

tetratricopeptide repeats 3<br />

Ifi49 15959 -4.5 -2.5 x x<br />

Ifitm3 interferon induced transmembrane protein 3<br />

Fgls, IP15, Cd225, mil-<br />

1, Cdw217<br />

66141 -5.0 -2.2 x x<br />

Irf7 interferon regulatory factor 7 54123 -7.3 -3.2 x x<br />

Isg20 interferon-stimulated protein DnaQL, HEM45 57444 -1.3 -3.9 x<br />

a Fold change values were calculated for the comparison <strong>of</strong> C57BL/6 BMM with the baseline <strong>of</strong> BALB/c BMM <strong>of</strong> mean values from<br />

three biological replicates per group. Positive values indicate higher expression in C57BL/6 BMM than in BALB/c BMM, whereas<br />

negative values indicate higher expression in BALB/c BMM than in C57BL/6 BMM. Differential expression between BMM <strong>of</strong> the two<br />

strains at the same treatment level (IFN-γ treated or medium control) using statistical testing in combination with an absolute fold<br />

change cut<strong>of</strong>f <strong>of</strong> 1.5 is indicated in bold.<br />

106


Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

Table R.3.15: Strain differences in gene expression <strong>of</strong> plasma membrane receptors between BMM <strong>of</strong> BALB/c and C57BL/6 mice.<br />

gene<br />

name<br />

description<br />

Rosetta Resolver Annotation fold change a IFN-γ effect<br />

alias<br />

Entrez<br />

Gene ID<br />

control IFN-γ BALB/c C57BL/6<br />

Cadm1 cell adhesion molecule 1<br />

Bl2, ST17, Igsf4, Necl2, Tslc1,<br />

Igsf4a, RA175A, RA175B, RA175C, 54725 -4.7 -4.0<br />

RA175N, SgIGSF, SynCam<br />

L1cam L1 cell adhesion molecule L1, CD171, L1-NCAM, NCAM-L1 16728 -2.2 -2.2<br />

Pcdh7 protocadherin 7 54216 -1.7 -1.7 x<br />

Itgal integrin alpha L Cd11a, LFA-1, Ly-15, Ly-21 16408 2.0 1.4 x x<br />

Cd14 CD14 antigen 12475 -3.2 -2.4<br />

Cd40 CD40 antigen<br />

IGM, p50, Bp50, GP39, IMD3,<br />

TRAP, HIGM1, T-BAM, Tnfrsf5<br />

21939 -1.2 -2.4 x x<br />

Fcgr1 Fc receptor, IgG, high affinity I CD64, IGGHAFC, FcgammaRI 14129 -2.0 -1.9 x x<br />

Tlr1 toll-like receptor 1 21897 -2.0 -1.6<br />

Procr protein C receptor, endothelial Ccca, EPCR 19124 2.3 2.3<br />

a Fold change values were calculated for the comparison <strong>of</strong> C57BL/6 BMM with the baseline <strong>of</strong> BALB/c BMM <strong>of</strong> mean values from<br />

three biological replicates per group. Positive values indicate higher expression in C57BL/6 BMM than in BALB/c BMM, whereas<br />

negative values indicate higher expression in BALB/c BMM than in C57BL/6 BMM. Differential expression between BMM <strong>of</strong> the two<br />

strains at the same treatment level (IFN-γ treated or medium control) using statistical testing in combination with an absolute fold<br />

change cut<strong>of</strong>f <strong>of</strong> 1.5 is indicated in bold.<br />

In the group <strong>of</strong> plasma membrane receptors, the adhesion molecules Cadm1, L1cam, and<br />

Pcdh7 were higher expressed in BALB/c BMM, whereas expression <strong>of</strong> integrin Itgal was higher in<br />

C57BL/6 BMM. The expression <strong>of</strong> CD14 / LPS receptor, CD40 / co-stimulatory receptor, Fcgr1 / high<br />

affinity IgG Fc receptor I, and Tlr1 / toll-like receptor 1, was higher in BALB/c BMM than in<br />

C57BL/6 BMM. Contrarily, Procr / protein C receptor exhibited higher expression in C57BL/6 BMM<br />

(Table R.3.15).<br />

Table R.3.16: Strain differences in gene expression <strong>of</strong> immune response related proteins, lysosomal, extracellular matrix degrading,<br />

and detoxification enzymes between BMM <strong>of</strong> BALB/c and C57BL/6 mice.<br />

Rosetta Resolver Annotation fold change a IFN-γ effect<br />

gene<br />

Entrez<br />

description<br />

alias<br />

name<br />

Gene ID<br />

control IFN-γ BALB/c C57BL/6<br />

Lbp lipopolysaccharide binding protein Ly88 16803 1.5 2.3<br />

Lta4h leukotriene A4 hydrolase 16993 1.9 1.7<br />

Oas2 2'-5' oligoadenylate synthetase 2 Oasl11 246728 -2.6 -2.5 x<br />

Oasl1 2'-5' oligoadenylate synthetase-like 1 oasl9 231655 -2.5 -4.9 x<br />

Arg2 arginase type II AII 11847 -5.1 -3.7<br />

Tfpi tissue factor pathway inhibitor EPI, LACI 21788 2.0 2.1<br />

Ctsc cathepsin C DPP1, DPPI 13032 1.5 1.7 x x<br />

Ctse cathepsin E CE, CatE 13034 -4.3 -5.7<br />

Ctsh cathepsin H 13036 2.7 2.0 x<br />

Hyal1 hyaluronoglucosaminidase 1 Hya1, Hyal-1 15586 -2.2 -1.7<br />

Mmp14 matrix metallopeptidase 14 (membrane-inserted)<br />

MT1-MMP, MT-<br />

MMP-1<br />

17387 2.3 1.8<br />

Gstm2 glutathione S-transferase, mu 2 Gstb2, Gstb-2 14863 3.5 3.3<br />

Gstp1 glutathione S-transferase, pi 1 GstpiB 14870 1.9 2.0<br />

a Fold change values were calculated for the comparison <strong>of</strong> C57BL/6 BMM with the baseline <strong>of</strong> BALB/c BMM <strong>of</strong> mean values from<br />

three biological replicates per group. Positive values indicate higher expression in C57BL/6 BMM than in BALB/c BMM, whereas<br />

negative values indicate higher expression in BALB/c BMM than in C57BL/6 BMM. Differential expression between BMM <strong>of</strong> the two<br />

strains at the same treatment level (IFN-γ treated or medium control) using statistical testing in combination with an absolute fold<br />

change cut<strong>of</strong>f <strong>of</strong> 1.5 is indicated in bold.<br />

LBP, LPS binding protein, and Lta4h, whose gene product catalyzes the first reaction step in<br />

the biosynthesis <strong>of</strong> leukotrienes, i. e. the reaction from leukotriene A4 to leukotriene B4, were<br />

higher expressed in C57BL/6 BMM than in BALB/c BMM. In contrast, 2'-5'-oligoadenylate<br />

synthetase genes Oas2 and Oasl1, which are part <strong>of</strong> the innate immune response, exhibited<br />

107


Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

higher expression in BALB/c BMM. The same direction <strong>of</strong> difference was observed for arginase<br />

type II (Arg2). Tissue factor pathway inhibitor (Tfpi), coding for a protease inhibitor with anticoagulant<br />

function, was higher expressed in C57BL/6 BMM. Genes <strong>of</strong> lysosomal, matrix<br />

degrading, and detoxification enzymes were differentially expressed between BMM <strong>of</strong> the two<br />

mouse strains. Lysosomal enzyme genes showed a mixed pattern with higher expression in<br />

BALB/c BMM for cathepsin E (Ctse) and hyaluronidas (Hyal1) and higher expression in C57BL/6<br />

BMM for cathepsins C and H (Ctsc, Ctsh). Matrix metallopeptidase 14 (Mmp14), which is involved<br />

in degradation <strong>of</strong> extracellular matrix, and glutathione S-transferases mu 1 and pi 1 (Gstm1,<br />

Gstp1) which are associated with detoxification processes, were detected with higher expression<br />

in C57BL/6 BMM (Table R.3.16).<br />

Strain-specific difference in expression <strong>of</strong> IFN-γ related genes<br />

BMM <strong>of</strong> BALB/c and C57BL/6 differ in their ability to eliminate infecting bacteria, especially<br />

after IFN-γ treatment (Breitbach et al. 2006). In this study, the primary reaction <strong>of</strong> BMM to IFN-γ<br />

in comparison between both strains was analyzed. Therefore, a network centered around the<br />

starting node IFN-γ was created using Ingenuity Pathway Analysis (IPA, www.ingenuity.com). This<br />

network contained 181 genes (Fig. R.3.5 A, B) additionally to the starting node IFN-γ. Genes were<br />

grouped manually into four categories: 1) difference existed between BALB/c BMM and C57BL/6<br />

BMM at both treatment levels, control and after IFN-γ treatment; 2) difference between BMM <strong>of</strong><br />

the two strains occurred only at control level; 3) difference between BMM <strong>of</strong> the two strains<br />

occurred only after IFN-γ treatment; 4) no strain difference was observed, but the gene<br />

expression was influenced at least in BMM <strong>of</strong> one mouse strain. A second categorization divided<br />

each group exhibiting strain differences into genes higher expressed in BALB/c BMM and into<br />

genes higher expressed in C57BL/6 BMM.<br />

After restricting the network to genes described to be linked to IFN-γ in macrophages or RAW<br />

cells in the Ingenuity Pathway Knowledge Base (IPKB), a list <strong>of</strong> 132 new macrophage-related<br />

genes out <strong>of</strong> 181 IFN-γ linked genes could be obtained. Correspondingly, a list <strong>of</strong> 49 <strong>of</strong> 181 genes<br />

that are already included in the IPKB as IFN-γ related in macrophages or RAW cells were<br />

exported.<br />

Fig. R.3.5: Ingenuity Pathway Analysis (IPA, www.ingenuity.com) <strong>of</strong> IFN-γ related genes.<br />

A, B. Network resulting from application <strong>of</strong> the grow-tool on transcriptome data<br />

To create a network centered around the starting node IFN-γ the grow-tool <strong>of</strong> IPA‘s pathway function was applied. The addition <strong>of</strong><br />

further nodes according to information stored in the so-called Ingenuity Pathway Knowledge Base (IPKB) was restricted to a list <strong>of</strong><br />

genes that are differentially expressed in at least 1 <strong>of</strong> 4 comparisons 1) IFN-γ effects in BALB/c BMM, 2) IFN-γ effects in C57BL/6 BMM,<br />

3) strain difference at non-treated control level and 4) strain difference after IFN-γ treatment. Further restrictions were not included.<br />

Networks are colored according to strain difference at control level (A) or at IFN-γ treated level (B). Green represents a higher<br />

expression in BALB/c BMM than in C57BL/6 BMM whereas red indicates higher expression in C57BL/6 BMM than in BALB/c BMM. The<br />

intensity <strong>of</strong> color correlates with the magnitude <strong>of</strong> difference: The shade <strong>of</strong> color becomes darker with increasing absolute fold<br />

change.<br />

108


Maren Depke<br />

Results<br />

Gene Expression Pattern <strong>of</strong> Bone-Marrow Derived Macrophages after Interferon-gamma Treatment<br />

A<br />

regulation <strong>by</strong> IFN-γ at least<br />

in BMM <strong>of</strong> one strain; no<br />

significant difference in<br />

expression between BMM<br />

<strong>of</strong> both strains<br />

difference between BMM <strong>of</strong><br />

the two mouse strains only<br />

after IFN-γ treatment<br />

B<br />

higher expression<br />

in BALB/c BMM<br />

higher expression<br />

in C57BL/6 BMM<br />

difference between BMM <strong>of</strong><br />

the two mouse strains only<br />

at control level<br />

difference between BMM <strong>of</strong><br />

the two mouse strains at<br />

both control and IFN-γ<br />

treated level<br />

regulation <strong>by</strong> IFN-γ at least<br />

in BMM <strong>of</strong> one strain; no<br />

significant difference in<br />

expression between BMM<br />

<strong>of</strong> both strains<br />

difference between BMM <strong>of</strong><br />

the two mouse strains only<br />

after IFN-γ treatment<br />

higher expression<br />

in BALB/c BMM<br />

higher expression<br />

in C57BL/6 BMM<br />

difference between BMM <strong>of</strong><br />

the two mouse strains only<br />

at control level<br />

difference between BMM <strong>of</strong><br />

the two mouse strains at<br />

both control and IFN-γ<br />

treated level<br />

109


cfu / S9 cell<br />

cfu / S9 cell<br />

% PI-negative S9 cells<br />

Maren Depke<br />

Results<br />

HOST CELL GENE EXPRESSION PATTERN IN AN<br />

IN VITRO INFECTION MODEL<br />

Staphylococcal cfu per <strong>host</strong> cell after internalization and <strong>host</strong> cell vitality [Melanie Gutjahr,<br />

Petra Hildebrandt]<br />

After 2.5 h, an averaged number <strong>of</strong> 1.3 staphylococcal colony forming units (cfu)/ S9 cell were<br />

found in internalization experiments (data courtesy <strong>of</strong> Melanie Gutjahr). Although not statistically<br />

significant in the four biological replicates used for this study, a trend <strong>of</strong> increasing cfu / S9 cell<br />

along with increasing infection time was discernible and led to an internalization mean value <strong>of</strong><br />

2.3 cfu / S9 cell 6.5 h after start <strong>of</strong> infection (Fig. R.4.1 A).<br />

A<br />

S9 infection proteome and transcriptome<br />

4<br />

3<br />

2<br />

1<br />

B<br />

100<br />

80<br />

60<br />

40<br />

20<br />

0<br />

RN1HG GFP RN1HG GFP<br />

infected S9 cells infected S9 cells<br />

2.5 h after start 6.5 h after start<br />

<strong>of</strong> time infection after start <strong>of</strong> infection <strong>of</strong> infection<br />

2.5 h<br />

6.5 h<br />

00<br />

untreated<br />

control<br />

1 2non-sorted3 4 5 sorted 6 7<br />

infected non-infected (GFP – ) infected (GFP + )<br />

2.5 h 6.5 h 2.5 h 6.5 h 2.5 h 6.5 h<br />

S9 cell samples<br />

Fig. R.4.1: Staphylococcal cfu per <strong>host</strong> cell after internalization and <strong>host</strong> cell vitality.<br />

A. Staphylococcal viable cell counts per S9 cell in infection experiments. The line indicates mean values per each analyzed time point.<br />

B. Percentages <strong>of</strong> viable PI – S9 cells in untreated control and after infection with S. aureus RN1HG GFP determined <strong>by</strong> FACS counting.<br />

Infected samples were analyzed for the two time points in different populations: without sorting or after sorting separately for GFP –<br />

and GFP + S9 cells. Mean values <strong>of</strong> n = 2 experiments.<br />

Cfu-data: courtesy <strong>of</strong> Melanie Gutjahr. FACS-data: courtesy <strong>of</strong> Petra Hildebrandt.<br />

Host cell vitality was determined <strong>by</strong> propidium iodide (PI) staining and counting in FACS<br />

(n = 2). Although 7.0 % and 12.1 % <strong>of</strong> cells in the infected, but non-sorted cell suspension were<br />

positive for PI and therefore not viable 2.5 h and 6.5 h after infection, respectively, compared to<br />

3.6 % <strong>of</strong> PI-positive dead cells in the untreated control, most <strong>of</strong> the dead, infected cells were<br />

discarded during the sorting procedure (Fig. R.4.1 B). Thus, cell vitality was not influenced in the<br />

subsets <strong>of</strong> sorted cells 2.5 h after infection (98.7 % in GFP + S9 cells, 99.0 % in GFP – S9 cells). After<br />

6.5 h, a minor drop in viability was recorded for the infected samples (97.3 % in GFP + S9 cells,<br />

98.8 % in GFP – S9 cells; data courtesy <strong>of</strong> Petra Hildebrandt).<br />

Reproducibility <strong>of</strong> replicates and clustering <strong>of</strong> treatment group members<br />

Staphylococcus aureus RN1HG GFP was used to infect S9 cells, a human bronchial epithelial<br />

cell line. The study included samples <strong>of</strong> two treatments, infected green fluorescence-positive<br />

FACS-sorted S9 cells and control cells that were exposed to medium, and analyzed the two time<br />

points 2.5 h and 6.5 h after start <strong>of</strong> infection. Hierarchical clustering visualized the grouping <strong>of</strong><br />

the array data sets according to their similarity (Fig. R.4.2). Biological replicates <strong>of</strong> each treatment<br />

111


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

and time point had the highest similarity to each other, and thus, the four treatment/time point<br />

groups defined the three lowest level clusters (Fig. R.4.2, orange dashed line). When further<br />

analyzing the next level <strong>of</strong> clustering the high similarity between the 2.5 h groups, i. e. infected<br />

samples and control samples, was obvious (Fig. R.4.2, red dashed line). In the following cluster<br />

level, the 6.5 h control samples showed higher similarity to both 2.5 h samples groups than to the<br />

6.5 h infected samples. Summing up, the biological replicates exhibited a high reproducibility. The<br />

highest similarity on the level <strong>of</strong> treatment groups was observed between 2.5 h infected and<br />

control samples, and not between the control samples <strong>of</strong> both time points as it might have been<br />

expected. The medium control samples <strong>of</strong> the 6.5 h timepoint were more similar to the 2.5 h<br />

samples (infection and control) than to the 6.5 h infected samples.<br />

medium control<br />

infected; GFP positive<br />

time point 2.5 h<br />

time point 6.5 h<br />

biological replicate 1<br />

biological replicate 2<br />

biological replicate 3<br />

biological replicate 4<br />

Fig. R.4.2: Hierarchical clustering <strong>of</strong> 16 array data sets from S9 infection experiment.<br />

The following clustering algorithms were applied on log-transformed data: Agglomerative clustering with average linkage using<br />

euclidian distance weighted <strong>by</strong> error as similarity measure. Sequences which were not expressed on all 16 arrays (defined <strong>by</strong> p > 0.01<br />

on intensity pr<strong>of</strong>ile level in Rosetta Resolver s<strong>of</strong>tware) and control sequences were excluded from the cluster analysis.<br />

The orange dashed line indicates the three lowest level clusters which led to a segmentation into the the four treatment/time point<br />

groups. The next level <strong>of</strong> clustering grouped the 6.5 h control samples next to 2.5 h control and infected samples (red dashed line).<br />

Comparison <strong>of</strong> treatment groups and assessment <strong>of</strong> differentially regulated genes<br />

For identification <strong>of</strong> differential gene expression, infected samples were compared to their<br />

time-matched controls. Additionally, same treatments were compared between the two time<br />

points for control purposes. In detail, four comparisons <strong>of</strong> sample groups were carried out:<br />

2.5 h infected GFP-positive samples vs. 2.5 h medium control<br />

6.5 h infected GFP-positive samples vs. 6.5 h medium control<br />

6.5 h infected GFP-positive samples vs. 2.5 h infected GFP-positive samples<br />

6.5 h medium control vs. 2.5 h medium control<br />

All comparisons are depicted <strong>by</strong> scatter plots (Fig. R.4.3). A small difference appeared<br />

between the infected and control samples 2.5 h after infection (Fig. R.4.3, panel ), while a<br />

strong reaction <strong>of</strong> infected cells compared to the time-matched control became visible after 6.5 h<br />

(Fig. R.4.3, panel ). Consequently, the infected samples strongly differed between both time<br />

points (Fig. R.4.3, panel ). But also in the control samples a certain change <strong>of</strong> signal intensities<br />

depending on incubation time was recorded. This change mainly consisted <strong>of</strong> lower signal values<br />

after 6.5 h than after 2.5 h (Fig. R.4.3, panel).<br />

112


infected; 2.5 h<br />

infected; 6.5 h<br />

mean intensity<br />

mean intensity<br />

infected; 6.5 h<br />

medium control; 6.5 h<br />

Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

<br />

<br />

medium control; 2.5 h<br />

infected; 2.5 h<br />

<br />

<br />

medium control; 6.5 h<br />

medium control; 2.5 h<br />

Fig. R.4.3: Scatter plots comparing mean signal intensities <strong>of</strong> treatment groups.<br />

The signals <strong>of</strong> the four groups “medium control sample after 2.5 h”, “GFP-positive sorted infected sample after 2.5 h”, “medium<br />

control sample after 6.5 h”, and “GFP-positive sorted infected sample after 2.5 h” were plotted after combining the four biological<br />

replicates. Control sequences and sequences that were absent on all <strong>of</strong> the arrays used for each scatter plot are not shown. (Absence<br />

<strong>of</strong> expression is defined via a p-value > 0.01 on intensity pr<strong>of</strong>ile level in Rosetta Resolver).<br />

When looking at examples <strong>of</strong> signal intensity data, gene expression pattern were recognized<br />

that were characterized <strong>by</strong> similar expression in the three groups <strong>of</strong> infected GFP-positive sorted<br />

cells after 2.5 h, medium control cells after 2.5 h, and infected GFP-positive sorted cells after<br />

6.5 h. A divergent signal intensity, lower or higher than in the three other groups, was noticeable<br />

for the last group <strong>of</strong> medium control cells after 6.5 h (Fig. R.4.4). However, the time-dependent<br />

changes in the control samples’ gene expression were not in the focus <strong>of</strong> this study, but the<br />

infection-dependent changes should be accessed. Therefore, a strategy was defined that allowed<br />

to exclude genes which are mainly influenced <strong>by</strong> a differing value in the group <strong>of</strong> 6.5 h medium<br />

control cells from the list <strong>of</strong> infection-dependent differential gene expression after 6.5 h <strong>of</strong><br />

infection.<br />

A<br />

Fkbp4 (EntrezGene ID 2288)<br />

B<br />

Cldn1 (EntrezGene ID 9076)<br />

2500<br />

3000<br />

Fig. R.4.4:<br />

Examples <strong>of</strong> mean signal<br />

intensities characterized <strong>by</strong> a<br />

higher (A) and lower value<br />

(B) for the medium control<br />

after 6.5 h compared to the<br />

other three sample groups.<br />

2000<br />

1500<br />

1000<br />

500<br />

0<br />

pMEM control<br />

tryps.<br />

medium ActD/NaN3<br />

control<br />

2.5 h 6.5 h<br />

2500<br />

2000<br />

1500<br />

1000<br />

500<br />

2288 FKBP4<br />

0<br />

pMEM control<br />

tryps.<br />

ActD/NaN3 medium<br />

control<br />

RN1HG infected pMEM control RN1HG infected<br />

RN1HG GFP wt GFP tryps. RN1HG GFP wt GFP<br />

infected,<br />

FACS sorted<br />

GFP positive<br />

GFPpositivpositive<br />

medium ActD/NaN3<br />

control<br />

infected,<br />

FACS sorted<br />

GFP positive<br />

GFP-<br />

2.5 h 6.5 h<br />

sorted<br />

sorted<br />

RN1HG infected pMEM control RN1HG infected<br />

RN1HG GFP wt GFP tryps. RN1HG GFP wt GFP<br />

FACS infected, sorted<br />

GFP positive<br />

GFPpositive<br />

ActD/NaN3 medium<br />

control<br />

FACS infected, sorted<br />

GFP positive<br />

GFP-<br />

6.5 h positive<br />

2.5 h<br />

sorted<br />

sorted<br />

2.5 h 6.5 h<br />

9076 CLDN1<br />

First, differential gene expression was determined with statistical testing for the comparison<br />

<strong>of</strong> infected GFP-positive sorted cells vs. medium control, each after 6.5 h. Second, the differential<br />

gene expression was calculated for the comparison <strong>of</strong> infected GFP-positive sorted cells after<br />

6.5 h with medium control cells after 2.5 h. Third, only genes which featured differential gene<br />

expression in both comparisons were included in the list <strong>of</strong> infection-dependent differential gene<br />

expression 6.5 h after infection. For assessing the infection-dependent differential gene<br />

113


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

expression after 2.5 h, infected GFP-positive sorted cells after 2.5 h were statistically compared<br />

to their time matched medium control cells.<br />

Altogether, 86 genes were defined as differentially expressed in infection compared to control<br />

2.5 h after infection. Of these genes, 40 possessed a minimal absolute fold change <strong>of</strong> 1.5, and 11<br />

genes held a minimal absolute fold change <strong>of</strong> 2 (Table R.4.1 A). In comparison, for the 6.5 h time<br />

point differential expression in infection compared to control was made up <strong>of</strong> 2296 genes, <strong>of</strong><br />

which 1196 passed an absolute fold change cut<strong>of</strong>f <strong>of</strong> 1.5 and still 495 passed a cut<strong>of</strong>f <strong>of</strong> 2<br />

(Table R.4.1 B).<br />

Comparison <strong>of</strong> infection-dependent differentially expressed genes after 2.5 h and 6.5 h<br />

When comparing the infection-dependent changes in gene expression with a minimal<br />

absolute fold change <strong>of</strong> 1.5 after 2.5 h and after 6.5 h, an overlap <strong>of</strong> 26 genes was recorded. This<br />

corresponds to 65 % <strong>of</strong> the genes differentially expressed 2.5 h after infection (Fig. R.4.5 A). If<br />

genes that did not exhibit differential expression relative to the baseline <strong>of</strong> 2.5 h medium control<br />

had not been excluded from the list <strong>of</strong> infection-dependent regulated genes after 6.5 h, the<br />

overlap would have even increased <strong>by</strong> 6 genes to 80 %.<br />

A<br />

B<br />

2.5 h<br />

repressed<br />

repressed 6.5 h<br />

14 26 1170<br />

induced<br />

5<br />

333<br />

induced<br />

9<br />

2<br />

837<br />

genes differentially expressed<br />

in S9 cells between infection<br />

with RN1HG GFP and medium<br />

control 2.5 h after infection<br />

genes differentially expressed<br />

in S9 cells between infection<br />

with RN1HG GFP and medium<br />

control 6.5 h after infection<br />

1<br />

23<br />

Fig. R.4.5: Comparison <strong>of</strong> infection-dependent regulated genes in S9 cells 2.5 h and 6.5 h after start <strong>of</strong> infection.<br />

After applying a minimal absolute fold change cut<strong>of</strong>f <strong>of</strong> 1.5, 40 genes were differentially expressed in infection compared to control<br />

2.5 h after start <strong>of</strong> infection, while the number <strong>of</strong> differentially expressed genes 6.5 h after start <strong>of</strong> infection increased to 1196. The<br />

overlap between both lists contained 26 genes (A). This group <strong>of</strong> 26 genes in the overlap could be subdivided into 23 genes which<br />

were induced 2.5 h as well as 6.5 h after start <strong>of</strong> infection, 2 genes which were repressed in both time points and 1 genes which was<br />

induced 2.5 h after start <strong>of</strong> infection and repressed after 6.5 h. In general, a bigger fraction <strong>of</strong> differentially regulated genes was<br />

induced than repressed in each time point. This summed up to 33 or 860 induced genes 2.5 h or 6.5 h after start <strong>of</strong> infection,<br />

respectively. Repression occurred for 7 genes at the 2.5 h time point and 336 genes at the 6.5 h time point (B).<br />

In a finer resolution, also the direction <strong>of</strong> regulation was comparable between the two<br />

analyzed time points. Of the 26 genes, which were differentially expressed after 2.5 h as well as<br />

after 6.5 h, 23 were induced and 2 were repressed in both time points. Only for one gene<br />

induction was observed after 2.5 h while the expression was repressed after 6.5 h (Fig. R.4.5 B).<br />

Concluding from this high degree <strong>of</strong> conformance in the two time points, the RNA expression<br />

pr<strong>of</strong>iling 2.5 h after start <strong>of</strong> infection seemed to catch the very early beginning <strong>of</strong> the <strong>host</strong> cell’s<br />

reaction, which was very strongly aggravated during the following 4 hours until the measurement<br />

6.5 h after start <strong>of</strong> infection.<br />

In each time point, induction <strong>of</strong> gene expression was more pronounced than repression.<br />

Induction summed up to 33 or 860 genes 2.5 h or 6.5 h after start <strong>of</strong> infection, respectively.<br />

Repression covered 7 genes at the 2.5 h time point and 336 genes at the 6.5 h time point<br />

(Fig. R.4.5 B).<br />

114


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

A Genes displaying statistically different expression values in the corresponding comparisons.<br />

testing a significant with p*


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

Differentially abundant proteins in infected samples compared to medium-treated controls<br />

[Melanie Gutjahr]<br />

Samples for transcriptome analysis were generated from the same experiments which yielded<br />

samples for a proteome study (Melanie Gutjahr). Here, a gel-free mass spectrometric analysis<br />

was performed, which allowed the identification <strong>of</strong> 1049 proteins in all analyzed samples. Of<br />

these, Melanie Gutjahr classified 112 and 224 as different in abundance 2.5 h and 6.5 h after start<br />

<strong>of</strong> infection, respectively. At the 2.5 h time point, reduced abundance prevailed with 73 proteins<br />

in comparison to 39 proteins with increased abundance. In contrary, 96 reduced and 128<br />

increased proteins were observed 6.5 h after start <strong>of</strong> infection. The common signature <strong>of</strong> both<br />

2.5 h and 6.5 h time point included 63 proteins, whereas 49 and 161 proteins were specifically<br />

regulated at the 2.5 h and 6.5 h time point, respectively. Reverse regulation at both time points<br />

did not exist.<br />

UniProt accession numbers <strong>of</strong> proteins exhibiting differential abundance were mapped to<br />

EntrezGene IDs using the UniProt ID mapping tool (www.uniprot.org). The majority <strong>of</strong> these<br />

EntrezGene IDs was available as annotation for sequences <strong>of</strong> the GeneChip Human Gene 1.0 ST<br />

array in Rosetta Resolver s<strong>of</strong>tware. Finally, 98 regulated proteins were available for comparison<br />

with transcriptome data from samples 2.5 h after infection, and at the 6.5 h time point, 198<br />

regulated proteins could be compared to array results (Table R.4.2).<br />

Table R.4.2: Mapping <strong>of</strong> UniProt entries to EntrezGene IDs and EntrezGene records in Rosetta Resolver s<strong>of</strong>tware.<br />

time<br />

point<br />

UniProt accession numbers<br />

<strong>of</strong> regulated proteins<br />

EntrezGene IDs<br />

after ID mapping<br />

EntrezGene IDs available in Rosetta Resolver s<strong>of</strong>tware<br />

for the GeneChip Human Gene 1.0 ST array<br />

2.5 h 112 110 98<br />

6.5 h 224 221 198<br />

Comparison <strong>of</strong> differentially abundant proteins with differentially expressed genes<br />

[in cooperation with Melanie Gutjahr]<br />

At the first time point 2.5 h after start <strong>of</strong> infection, the small list <strong>of</strong> differentially expressed<br />

genes did not overlap with the list <strong>of</strong> regulated proteins.<br />

When comparing transcriptome and proteome results <strong>of</strong> the 6.5 h time point, an overlap <strong>of</strong> 11<br />

genes/proteins was discernible. These included 7 (ALCAM, APOL2, ERAP1, IFIT2, IFIT3, OASL,<br />

WARS) and 3 (FASN, KPNA2, PPME1) genes/proteins which exhibited in both analysis methods<br />

up- and down-regulation, respectively. The remaining gene/protein DYNC1H1 was induced at<br />

transcript level and reduced in protein abundance.<br />

The comparison between differentially expressed genes at the earlier 2.5 h time point with<br />

regulated proteins at the later 6.5 h time point revealed that for 2 (IFIT2, IFIT3) increased<br />

proteins the transcripts already were induced 2.5 h after start <strong>of</strong> infection.<br />

When the comparison was performed in reversion, i. e. when the 2.5 h time point regulated<br />

proteins were compared with the later 6.5 h time point’s gene expression changes, 4 (CNP,<br />

DYNC1H1, IFIT1, ZC3HAV1) proteins were reduced at 2.5 h whereas gene transcripts were<br />

induced at 6.5 h. Furthermore, one protein (ALCAM) was increased after 2.5 h, but the transcript<br />

was induced only at the 6.5 h time point. This gene/protein was already included in a list<br />

mentioned above. Finally, the protein FASN was reduced 2.5 h after start <strong>of</strong> infection, but the<br />

transcript was repressed after 6.5 h. The FASN gene/protein was already included in the list <strong>of</strong><br />

gene and protein repression at the 6.5 h time point (Table R.4.3).<br />

116


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

Table R.4.3: Differentially abundant proteins which exhibit differential gene expression.<br />

gene name<br />

description<br />

UniProt<br />

accession<br />

number<br />

EntrezGene<br />

ID<br />

protein<br />

fold change a<br />

transcript<br />

2.5 h 6.5 h 2.5 h 6.5 h<br />

ALCAM activated leukocyte cell adhesion molecule Q13740 214 4.5 4.3 1.1 1.8<br />

APOL2 apolipoprotein L, 2 Q9BQE5 23780 2.1 4.5 1.2 4.5<br />

ERAP1 endoplasmic reticulum aminopeptidase 1 Q9NZ08 51752 N/A 5.6 -1.0 1.6<br />

IFIT2<br />

interferon-induced protein with<br />

tetratricopeptide repeats 2<br />

P09913 3433 1.1 6.2 2.4 5.2<br />

IFIT3<br />

interferon-induced protein with<br />

tetratricopeptide repeats 3<br />

O14879 3437 1.1 2.6 1.6 3.4<br />

OASL 2'-5'-oligoadenylate synthetase-like Q15646 8638 N/A 16.9 1.1 5.1<br />

WARS tryptophanyl-tRNA synthetase P23381 7453 -1.3 1.6 1.1 4.4<br />

FASN fatty acid synthase P49327 2194 -1.6 -1.6 -1.1 -1.7<br />

KPNA2<br />

karyopherin alpha 2 (RAG cohort 1, importin<br />

alpha 1)<br />

P52292 3838 -1.5 -1.7 -1.3 -1.7<br />

PPME1 protein phosphatase methylesterase 1 Q9Y570 51400 -2.9 -18.0 -1.1 -1.9<br />

CNP 2',3'-cyclic nucleotide 3' phosphodiesterase P09543 1267 -2.4 1.2 -1.1 1.6<br />

IFIT1<br />

interferon-induced protein with<br />

tetratricopeptide repeats 1<br />

P09914 3434 -2.1 1.3 1.6 3.3<br />

ZC3HAV1 zinc finger CCCH-type, antiviral 1 Q7Z2W4 56829 -2.0 1.2 1.4 3.4<br />

a Fold change values were calculated for the comparison <strong>of</strong> infected GFP + S9 cell with the baseline <strong>of</strong> medium control samples.<br />

Differential regulation is indicated in bold.<br />

Evaluation <strong>of</strong> infection-dependent differentially expressed genes using Ingenuity Pathway<br />

Analysis (IPA) for the two time points 2.5 h and 6.5 h after start <strong>of</strong> infection<br />

For each <strong>of</strong> the two analyzed time points, 2.5 h and 6.5 h after start <strong>of</strong> infection, a list with<br />

genes differentially expressed in infected samples compared to controls was uploaded to<br />

Ingenuity Pathway Analysis (IPA, www.ingenuity.com). Adequate fold change value accompanied<br />

the EntrezGene identifiers. A minimal absolute fold change cut<strong>of</strong>f <strong>of</strong> 1.5 had been applied.<br />

In the statistical testing performed <strong>by</strong> the different IPA tools, the size <strong>of</strong> the input list<br />

influences the significance levels <strong>of</strong> the results. As the two lists were very different in size (40<br />

genes at the 2.5 h time point vs. 1196 genes at the 6.5 h time point; Table R.4.1), the significance<br />

levels for each analysis were <strong>by</strong> far not directly comparable. Therefore, the 2.5 h data were<br />

examined separately and additionally served for illustration purposes like overlaying gene<br />

expression data for 2.5 h on 6.5 h data results. The analysis concerning networks and pathways<br />

was focused on the differential gene expression at the 6.5 h time point.<br />

In the samples <strong>of</strong> the time point 2.5 h after infection, the big fraction <strong>of</strong> genes with increased<br />

expression was noticeable (33 <strong>of</strong> 40 genes; Fig. R.4.5 B). Strongest induction (fold change 4.2)<br />

was observed for IL-6, a cytokine occupying a key position in regulation <strong>of</strong> the immune response.<br />

Another immune-stimulating cytokine, IFN-β, was strongly induced (IFNB1, fold change 2.8;<br />

rank 4 <strong>of</strong> induced genes), and the response to interferon was seen in the induction <strong>of</strong> interferoninduced<br />

proteins IFIT2 (fold change 2.4) and IFIT3 (fold change 1.6). IFN-β also influences IL-6<br />

expression (Fig. R.4.6).<br />

Interestingly, prostaglandin-endoperoxide synthase 2 transcription was induced (PTGS2 alias<br />

Cox-2, fold change 2.3; rank 6 <strong>of</strong> induced genes). This enzyme generates the inflammation<br />

mediators prostaglandin (PG) G 2 and H 2 from arachidonic acid. From PGH 2 , the other<br />

prostaglandins like PGE 2 are formed <strong>by</strong> different synthases. In the data set, also PGE 2 receptor<br />

PTGER4 was induced (and induction <strong>of</strong> PGE 2 receptor PTGER2 was seen only at the 6.5 h time<br />

point). Therefore, prostaglandin synthesis as well as prostaglandin receptor were induced on<br />

117


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

transcriptome level at the 2.5 h time point. PTGS2 transcription is positively influenced <strong>by</strong><br />

IL-6. All three genes, IL6, PTGS2, and PTGER4, are indirectly induced <strong>by</strong> endothelin, EDN1, which<br />

also has vasoconstrictor functions and itself was induced 2.5 h after start <strong>of</strong> infection (fold<br />

change 4.2). The influence on expression is at least in parts mediated <strong>by</strong> ERK1/2, which was not<br />

regulated itself (Fig. R.4.6).<br />

Fig. R.4.6:<br />

Interactions <strong>of</strong> IL-6, PTGS2,<br />

PTGER4, EDN1, ERK1/2, IFNB1,<br />

IFIT2, and IFIT3 in a custom<br />

pathway design (modified from<br />

IPA, www.ingenuity.com).<br />

Red color indicates induction <strong>of</strong><br />

gene expression in infected<br />

samples compared to control 2.5 h<br />

after start <strong>of</strong> infection. More<br />

intense shade <strong>of</strong> color is used for<br />

higher absolute fold change<br />

values. Dashed lines indicate<br />

indirect influence on expression<br />

(E), activation (A), phosphorylation<br />

(P), and secretion (S), the solid line<br />

binding (B).<br />

These genes illustrated a beginning pro-inflammatory response. Another interesting candidate<br />

which is assigned an important role at the interface <strong>of</strong> neuronal and immune system, the<br />

leukemia inhibitory factor LIF, was increased in infected cells (fold change 1.6) already 2.5 h after<br />

start <strong>of</strong> infection (and stayed induced at the 6.5 h time point). LIF belongs to the IL-6 family <strong>of</strong><br />

cytokines and recruits, like IL-6, the ubiquitous glycoprotein gp130 as second subunit <strong>of</strong> its<br />

receptor.<br />

Several genes for signal transduction molecules and transcription factors were included in the<br />

small set <strong>of</strong> 40 differentially expressed genes 2.5 h after start <strong>of</strong> infection. An increased<br />

expression was observed for NR4A2 (fold change 3.5), NFKBIZ (3.0), KLF4 (2.1), ATF3 (2.1), BCL6<br />

(1.7), MYC (1.7), ZFP36L2 (1.6), and KLF6 (1.5).<br />

Increased expression was recorded for TNFAIP3 (2.3), a gene thought to be critical for limiting<br />

inflammation, and for ZC3H12C (1.5), a member <strong>of</strong> a family <strong>of</strong> negative regulators in macrophage<br />

activation. In addition, CD274 (B7-H1), ligand for the immunoinhibitory receptor PD-1 on<br />

activated B and T cells and on monocytes, was higher expressed in infected samples than in<br />

controls. These gene expression changes alluded to a role in counterregulatory processes during<br />

the beginning <strong>of</strong> infection and inflammation.<br />

Other genes like NEDD9 (2.0), TNC (1.6), and RND3 (1.5) are involved in cell adhesion,<br />

rounding and cytoskeleton organization, indicating the beginning <strong>of</strong> morphological changes in the<br />

infected cells.<br />

Even more, cell cycle progression seems to be negatively affected <strong>by</strong> the infection, as seen <strong>by</strong><br />

increased transcription <strong>of</strong> KLF4 (2.1), PPP1R15A (2.1), and MYC (1.7). Strikingly, 7 <strong>of</strong> 40 genes<br />

were repressed 2.5 h after start <strong>of</strong> infection and all <strong>of</strong> them are functionally associated with cell<br />

cycle progression: AURKA (−1.9), PLK1 (−1.8), HIST1H2AB (−1.6), BUB1 (−1.5) C13ORF34/BORA<br />

(−1.5), CDCA2 (−1.5), and CDCA8 (−1.5).<br />

About two thirds <strong>of</strong> regulated genes in the 2.5 h samples were still regulated 6.5 h after<br />

infection, and the absolute fold change was even increased for almost all <strong>of</strong> these genes at the<br />

later time point (Fig. R.4.7). Only one exception with reversed direction <strong>of</strong> regulation occurred,<br />

and PTGS2 had a smaller fold change after 6.5 h than after 2.5 h although its expression still was<br />

increased. To conclude, the changes started in the early time point were enforced in the later<br />

time point.<br />

118


fold change (infected vs. control)<br />

fold change (infected vs. control)<br />

fold change (infected vs. control)<br />

Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

14<br />

12<br />

14<br />

12<br />

14<br />

12<br />

Fig. R.4.7:<br />

2<br />

Comparison <strong>of</strong> fold change value for differentially expressed genes both after<br />

2.5 h and after 6.5 h.<br />

0<br />

Absolute fold change values for almost all <strong>of</strong> these genes increased at the later<br />

-2<br />

time point (black lines). PTGS2 had a smaller fold change after 6.5 h than after<br />

2.5 h (gray dashed line) and only one gene, ADAMTS1, showed an reversed -4<br />

direction <strong>of</strong> regulation in both time points (gray dashed/dotted line).<br />

10<br />

8<br />

6<br />

4<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

2.5 h 2.5 h 2.5 h 6.5 h 6.5 h 6.5 h<br />

In IPA, a global functional analysis can be applied to the complete data set. Results <strong>of</strong>fer an<br />

overview on the data-associated biological functions with a p-value to judge on overrepresentation<br />

<strong>of</strong> the corresponding function in the data set. The results can be grouped in three<br />

topics: 1) Diseases and Disorders 2) Molecular and Cellular Functions, and 3) Physiological System<br />

Development and Function. Here, especially the first and second topic were relevant for data<br />

analysis because the data were generated in vitro. To get a first impression, only the first five<br />

categories with smallest p-value were selected from the topics 1) and 2) for the 6.5 h time point.<br />

In the topic “Diseases and Disorders” a very dominant association <strong>of</strong> the data set with diverse<br />

inflammation and infection related categories was visible (Table R.4.4). Category “Cancer”<br />

obtained the smallest p-value, but as such category collects diverse functions that might also be<br />

immune-system associated, the result will not be interpreted as cancer like reaction <strong>of</strong> S9 cells to<br />

infection. Indeed, in a comparison <strong>of</strong> all genes in the list associated with “Cancer” and all genes<br />

linked to the other four immune-related categories in this analysis, approximately one third<br />

appeared in both “Cancer” as well as in the immune-related categories.<br />

The top categories <strong>of</strong> “Molecular and Cellular Function” mainly dealt with cellular growth,<br />

proliferation, death or related aspects (Table R.4.4). First signs <strong>of</strong> the influence <strong>of</strong> infection on<br />

this issue have already been recognized in the 2.5 h samples. In the later analysis time point <strong>of</strong><br />

6.5 h this influence was even more manifested and resulted in very low p-values.<br />

Table R.4.4: Global functional analysis (IPA, www.ingenuity.com) <strong>of</strong> the infection-dependent differentially regulated genes in S9 cells<br />

6.5 h after start <strong>of</strong> infection. First five categories <strong>of</strong> the topics “Diseases and Disorders” and “Molecular and Cellular Functions” are<br />

cited with the range <strong>of</strong> p-values <strong>of</strong> the associated sub-categories.<br />

rank<br />

Diseases and Disorders<br />

Molecular and Cellular Functions<br />

category p-value range category p-value range<br />

1 Cancer 3.44E-09 – 1.85E-03 Cellular Growth and Proliferation 1.04E-12 – 2.11E-03<br />

2 Infection Mechanism 3.76E-09 – 1.50E-03 Cell Death 1.27E-10 – 2.21E-03<br />

3 Antimicrobial Response 5.46E-07 – 7.96E-04 Cellular Development 4.94E-09 – 2.06E-03<br />

4 Inflammatory Response 5.46E-07 – 1.84E-03 Cellular Function and Maintenance 8.19E-08 – 1.95E-03<br />

5 Immunological Disease 1.12E-06 – 1.76E-03 Cell Cycle 1.22E-07 – 2.21E-03<br />

The Canonical Pathways tool in IPA allows analysis <strong>of</strong> the data set in the context <strong>of</strong> predefined<br />

pathways and their associated genes, which are displayed in order <strong>of</strong> cellular location and<br />

function and which can be overlaid with gene expression data.<br />

The first, most significant result for the infection-dependent gene expression changes in S9<br />

cells 6.5 h after start <strong>of</strong> infection was the pathway “Interferon Signaling” (p = 5.81E−09;<br />

ratio = 13/30; Fig. R.4.8). Induced gene expression was documented for IFN-β, the signal<br />

transduction molecules JAK2, STAT1, and STAT2, the regulatory suppressor <strong>of</strong> cytokine signaling<br />

119


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

SOCS1, and finally for the transcriptionally regulated genes PSMB8, TAP1, IRF1, IFI35, IFIT1, IFIT3,<br />

OAS1, and MX1.<br />

For the influence <strong>of</strong> staphylococcal infection on IFN signaling in in vivo kidney gene expression<br />

please refer to Results / Kidney Gene Expression Pattern in an in vivo Infection Model / Fig. R.2.9,<br />

page 86.<br />

Fig. R.4.8:<br />

Interferon Signaling (modified<br />

from IPA, www.ingenuity.com).<br />

Red color indicates increase <strong>of</strong><br />

expression. More intense shade<br />

<strong>of</strong> color is used for higher<br />

absolute fold change values.<br />

The second result in significantly over-represented canonical pathways was “Role <strong>of</strong> Pattern<br />

Recognition Receptors in Recognition <strong>of</strong> Bacteria and Viruses” (p = 3.63E−08; ratio = 20/80). This<br />

pathway includes C3AR1 (fold change 3.6), receptor for complement component C3a, pentraxin<br />

PTX3 (3.7), a soluble pattern recognition receptor (PRR) and inflammation regulator involved in<br />

enhancing complement activation and clearance <strong>of</strong> apoptotic cells, toll-like receptor TLR3 (5.1)<br />

with its signal transduction molecule MYD88 (2.3), intracellular PRR RIG-1 (DDX58; 3.2) and MDA-<br />

5 (IFIH1; 5.6), which are involved in viral double-stranded RNA recognition, and NOD1 (2.5),<br />

receptor for bacterial diaminopimelic acid with its signal transduction kinase RIPK2 (3.8). NOD1 is<br />

known to activate CASP1 (4.4), a protease cleaving the inactive IL-1β precursor.<br />

Three different 2'-5'-oligoadenylate synthetase genes OAS1/2/3 (2.5/2.4/1.9) were induced.<br />

They are part <strong>of</strong> the innate immune response, induced <strong>by</strong> interferon, and their product<br />

oligoadenylate activates RNASEL (1.6), which was additionally induced on mRNA level in this<br />

experiment. Cytokine genes IL6 (11.9), IFNB1 (3.6), IL12A (4.0), CCL5 (4.0; RANTES) are included in<br />

this pathway as end point <strong>of</strong> signal transduction starting with different TLRs via NFκB. Finally,<br />

three subunits <strong>of</strong> PI3K, phosphoinositide-3-kinase, were induced: PIK3R1 (1.7), regulatory subunit<br />

1/alpha, PIK3R3 (2.4), regulatory subunit 3/gamma, and PIK3CA (1.7), catalytic subunit alpha<br />

polypeptide.<br />

Significant over-representation was also observed for members <strong>of</strong> the pathway “Death<br />

Receptor Signaling” (p = 1.21E−03; ratio = 12/64). The death receptor FAS (3.7) and cascade<br />

initiating caspases CASP8 (1.6) and CASP10 (2.2) were induced, but also CFLAR (2.6; FLIP) as<br />

inhibitor <strong>of</strong> CASP8/10. Ligands TNFSF10 (18.5; APO2L, TRAIL) and TNFSF15 (4.6; TL1) were<br />

strongly induced, but not their receptors (DR). Death receptor’s/TNF receptor’s signal<br />

transduction kinase RIPK1 (2.2) was induced together with MAP kinases MAP3K14 (2.0) and<br />

MAP4K4 (2.7) and IκB kinase (IKK) subunit IKBKE (2.3). Additionally, CASP3 was slightly induced<br />

120


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

(1.6), but CASP9, the link to the intrinsic mitochondrial apoptosis pathway, was repressed (−1.6).<br />

Possibly, S9 cells were prepared to react to external apoptosis signals or give signals to infection<br />

relevant immune cells like neutrophils or macrophages. Of course, such situation does not occur<br />

in the S9 cell in vitro infection setting.<br />

Similarly, antigen presentation will not find any reaction partner in vitro. This canonical<br />

pathway was significantly overrepresented in the infection-dependent genes which were<br />

differentially expressed 6.5 h after start <strong>of</strong> infection (p = 1.21E−02; ratio = 8/39; Fig. R.4.9).<br />

Antigen presentation is divided into two parts, presentation <strong>of</strong> intracellular and extracellular<br />

antigens. Intracellular antigens are digested <strong>by</strong> cytosolic proteasome with sequential transport <strong>of</strong><br />

peptides into the endoplasmatic reticulum (ER). These peptides are loaded to MHC-I molecules<br />

and transported via the Golgi to the cell surface for presentation. The immune-proteasome gene<br />

PSMB8 (LMP7), coding for a protein in the 20S-core <strong>of</strong> the proteasome, was induced (fold change<br />

1.8). PSMB9 (LMP2) with a fold change <strong>of</strong> 1.499 was just marginally below the cut<strong>of</strong>f and could<br />

be regarded as induced, too. Both induced genes lead to an immune response-specific<br />

reorganization <strong>of</strong> the proteasomes’ catalytic center. Induced were the peptide transporters TAP1<br />

(2.4) and TAP2 (2.2), the TAP-binding protein tapasin/TPN (TAPBP; 1.5), which attaches the TAP<br />

molecules to the MHC-I complex, and the ER aminopeptidases ERAP1 (1.6) and ERAP2 (4.4),<br />

which are responsible for N-terminal trimming <strong>of</strong> the peptide in the ER. At the end <strong>of</strong> this<br />

pathway, MHC-I molecules HLA-E (1.7) and HLA-F (1.8) and MHC class I-related molecule MR1<br />

(2.0) were induced. MR1, a non-classical MHC-gene, features a strong homology to MHC-I<br />

molecules, is highly conserved in human and mouse, ubiquitously transcribed in different tissue<br />

or cell types, and speculated to present an unknown invariant ligand to certain “innate” mucosaassociated<br />

invariant T cells (MAIT cells) expressing an invariant T cell receptor (TCR) similar as<br />

innate natural killer T cells (iNKT) (Riegert et al. 1998, Hansen TH et al. 2007). Beta-2-<br />

microglobulin (B2M; MHC-I-β) as the second subunit <strong>of</strong> the MHC-I complex showed a trend <strong>of</strong><br />

induction with a fold change <strong>of</strong> 1.3, but was not significant in statistical testing.<br />

Fig. R.4.9: Antigen Presentation (modified from IPA, www.ingenuity.com). Red color indicates increase <strong>of</strong> expression.<br />

121


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

Immune cells and other cells <strong>of</strong> the body are capable <strong>of</strong> presenting intracellular antigens via<br />

MHC-I. Although a significant increase in expression was observed for MHC-II gene HLA-DOP<br />

(3.8), all other MHC-II genes for both alpha and beta subunits were not differentially expressed.<br />

Additionally, the intensity <strong>of</strong> MHC-II gene expression is clearly among the lower values on the<br />

whole array. This is not surprising, because presentation <strong>of</strong> extracellular antigens via MHC-II<br />

preferentially occurs <strong>by</strong> pr<strong>of</strong>essional antigen presenting cells (APC) like dendritic cells (DC),<br />

macrophages and B cells, and the human bronchial epithelial cell line S9, which was used in this<br />

study, does not belong to the group <strong>of</strong> pr<strong>of</strong>essional APC.<br />

For the influence <strong>of</strong> staphylococcal infection on antigen presentation in in vivo kidney gene<br />

expression please refer to Results / Kidney Gene Expression Pattern in an in vivo Infection<br />

Model / Fig. R.2.12, page 88.<br />

Persistence and enhancement <strong>of</strong> S9 reaction to infection with S. aureus RN1HG from the 2.5 h<br />

to the 6.5 h time point<br />

It has already been mentioned that a big fraction <strong>of</strong> the genes, which were differentially<br />

expressed at the 2.5 h time point, was still regulated at the later 6.5 h time point. This concerned<br />

26 genes (Table R.4.5).<br />

Table R.4.5: Overview on genes which were differentially expressed in S9 cells at both the 2.5 h and the 6.5 h time point in<br />

comparison between infection with S. aureus RN1HG GFP and control treatment.<br />

Rosetta Resolver annotation<br />

fold change a<br />

gene name<br />

description<br />

Entrez<br />

Gene ID<br />

2.5 h 6.5 h<br />

IL6 interleukin 6 (interferon, beta 2) 3569 4.2 11.9<br />

NFKBIZ nuclear factor <strong>of</strong> kappa light polypeptide gene enhancer in B-cells inhibitor, zeta 64332 3.0 7.9<br />

IFNB1 interferon, beta 1, fibroblast 3456 2.8 3.6<br />

IFIT2 interferon-induced protein with tetratricopeptide repeats 2 3433 2.4 5.2<br />

PTGS2 prostaglandin-endoperoxide synthase 2 (prostaglandin G/H synthase and cyclooxygenase) 5743 2.3 1.6<br />

KLF4 Kruppel-like factor 4 (gut) 9314 2.1 9.9<br />

ATF3 activating transcription factor 3 467 2.1 3.9<br />

PPP1R15A protein phosphatase 1, regulatory (inhibitor) subunit 15A 23645 2.1 2.4<br />

EDN1 endothelin 1 1906 2.0 8.2<br />

NEDD9 neural precursor cell expressed, developmentally down-regulated 9 4739 2.0 5.5<br />

PMAIP1 phorbol-12-myristate-13-acetate-induced protein 1 5366 2.0 4.4<br />

PTGER4 prostaglandin E receptor 4 (subtype EP4) 5734 1.8 5.2<br />

ADAMTS1 ADAM metallopeptidase with thrombospondin type 1 motif, 1 9510 1.8 -1.7<br />

BCL6 B-cell CLL/lymphoma 6 604 1.7 3.5<br />

SAT1 spermidine/spermine N1-acetyltransferase 1 6303 1.6 4.6<br />

ZFP36L2 zinc finger protein 36, C3H type-like 2 678 1.6 4.0<br />

FAM46A family with sequence similarity 46, member A 55603 1.6 3.5<br />

IFIT3 interferon-induced protein with tetratricopeptide repeats 3 3437 1.6 3.4<br />

TNC tenascin C 3371 1.6 2.1<br />

LIF leukemia inhibitory factor (cholinergic differentiation factor) 3976 1.6 2.2<br />

CD274 CD274 molecule 29126 1.6 5.5<br />

KLF6 Kruppel-like factor 6 1316 1.5 2.6<br />

ZC3H12C zinc finger CCCH-type containing 12C 85463 1.5 2.8<br />

C6orf141 chromosome 6 open reading frame 141 135398 1.5 3.7<br />

CDCA8 cell division cycle associated 8 55143 -1.5 -2.1<br />

HIST1H2AB histone cluster 1, H2ab 8335 -1.6 -2.1<br />

a Fold change values were calculated for the comparison <strong>of</strong> infected GFP + S9 cell with the baseline <strong>of</strong> medium control samples.<br />

Leukemia inhibitory factor (LIF) was already increased 2.5 h after start <strong>of</strong> infection (fold<br />

change 1.6). This increase was further elevated after 6.5 h (2.2). Fittingly, increased expression<br />

was also detectable for the LIF receptor (LIFR, 2.2) at this time point.<br />

122


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

CD274 alias programmed cell death 1 ligand 1 (PD-L1) is involved in the regulation <strong>of</strong><br />

inflammatory responses (Sharpe et al. 2007). In this study, CD274 gene expression was induced<br />

2.5 h after start <strong>of</strong> infection and further increased at the 6.5 h time point. Interestingly, at the<br />

later time point, the second PD-1 receptor, programmed cell death 1 ligand 2 (PDCD1LG2; 2.7),<br />

was also induced.<br />

The prostaglandin-endoperoxide synthase 2 (PTGS2) was less induced at the 6.5 h than at the<br />

2.5 h time point (Table R.4.5). Additionally, the intensity values for PTGS2 at 6.5 h were lower in<br />

both control and infected samples (60 and 94) when compared to the 2.5 h control sample (280).<br />

This might hint for an early reaction that was about to stop at the later time point. Induction <strong>of</strong><br />

prostaglandin receptor PTGER4 still increased from 2.5 h to 6.5 h, and induction <strong>of</strong> another<br />

receptor, PTGER2, was only detectable at the 6.5 h time point. The reaction <strong>of</strong> the genes for<br />

these receptors seemed to lag behind the reaction <strong>of</strong> the gene coding for the enzyme responsible<br />

for synthesis <strong>of</strong> their ligand.<br />

Interferon-β (IFNB1) was induced at both analyzed time points. At the 2.5 h time point,<br />

interferon reaction has already been described in reference to the induced genes IFIT2 and IFIT3<br />

(page 117). In the IPA canonical pathway “Interferon Signaling” (page 119), IFI35, IFIT1, IFIT3,<br />

IRF1, and others were included as induced, interferon-regulated genes at the 6.5 h time point,<br />

and IFIH1 has already been introduced in the IPA canonical pathway “Role <strong>of</strong> Pattern Recognition<br />

Receptors in Recognition <strong>of</strong> Bacteria and Viruses” (page 120). Manual filtering revealed further<br />

six interferon regulated genes IFI16, IFI44L, IFIT2, IFIT5, IRF2, and ISG20 to be induced 6.5 h after<br />

start <strong>of</strong> infection.<br />

A histone cluster 1 gene (HIST1H2AB) was repressed significantly with a less than −1.5x fold<br />

change 2.5 h after start <strong>of</strong> infection. But 6.5 h after infection, a number <strong>of</strong> 17 other histone<br />

cluster 1 genes together with HIST1H2AB were repressed. Of these 17 additional genes, 4 had<br />

already yielded a significant p-value at the 2.5 h time point, but did not pass the 1.5 fold cut<strong>of</strong>f.<br />

Furthermore, H1 histone family member 0 (H1F0) and H2A histone family member X (H2AFX)<br />

were repressed at the 6.5 h time point (Table R.4.5, Table R.4.6).<br />

Cyclins CCNB1, CCNB2, and CCNH were repressed, while CCND1 and CCNL1 were induced<br />

6.5 h after start <strong>of</strong> infection. In this context, induced cyclin-dependent kinases / kinase-like CDK6,<br />

CDKL2, and CDKL5 were conspicuous at the 6.5 h time point, while cyclin-dependent kinase<br />

inhibitors CDKN1B, CDKN2B, and CDKN3 were repressed. Contrarily, the inhibitor CDKN2C was<br />

induced. Similarly, genes coding for cell division cycle associated proteins CDCA3, CDCA4, CDCA8,<br />

and CDC25A, and genes encoding centromer/centrosomal proteins CENPF, CEP55, CEP97,<br />

CEP120, and CEP250 were repressed, while CDC14A and CEP170 were induced. Regulatory genes<br />

CKS2, GAS2L3, PRC1, and SKP2 were repressed, whereas BTG1, CHEK2, and DMTF1 were<br />

observed as induced (Table R.4.6).<br />

In a condensed manner, histones were repressed, some cyclin-dependent kinase (CDK) were<br />

induced, and CDK inhibitors, cyclins, cell division cycle associated proteins, centromer/<br />

centrosomal proteins and regulatory genes were found with both repressed and induced<br />

examples. Nevertheless, repression held a bigger fraction <strong>of</strong> this subset <strong>of</strong> genes than induction.<br />

123


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

Table R.4.6: Overview on selected differentially expressed cell cycle and proliferation related genes in S9 cells 6.5 h after start <strong>of</strong><br />

infection with S. aureus RN1HG.<br />

Rosetta Resolver annotation<br />

fold change a<br />

gene name<br />

description<br />

Entrez Gene<br />

ID<br />

2.5 h 6.5 h<br />

HIST1H1A histone cluster 1, H1a 3024 -1.4 -3.0<br />

HIST1H1B histone cluster 1, H1b 3009 -1.4 -1.7<br />

HIST1H1C histone cluster 1, H1c 3006 -1.3 -1.9<br />

HIST1H1D histone cluster 1, H1d 3007 -1.4 -2.2<br />

HIST1H2AB histone cluster 1, H2ab 8335 -1.6 -2.1<br />

HIST1H2AC histone cluster 1, H2ac 8334 -1.4 -1.6<br />

HIST1H2AI histone cluster 1, H2ai 8329 -1.3 -1.7<br />

HIST1H2AK histone cluster 1, H2ak 8330 -1.4 -2.2<br />

HIST1H2BH histone cluster 1, H2bh 8345 -1.3 -1.8<br />

HIST1H2BN histone cluster 1, H2bn 8341 -1.3 -1.5<br />

HIST1H3A histone cluster 1, H3a 8350 -1.5 -2.1<br />

HIST1H3B histone cluster 1, H3b 8358 -1.3 -1.5<br />

HIST1H3F histone cluster 1, H3f 8968 -1.3 -1.8<br />

HIST1H3I histone cluster 1, H3i 8354 -1.3 -1.7<br />

HIST1H4B histone cluster 1, H4b 8366 -1.4 -2.0<br />

HIST1H4C histone cluster 1, H4c 8364 -1.4 -1.6<br />

HIST1H4D histone cluster 1, H4d 8360 -1.5 -1.7<br />

HIST1H4F histone cluster 1, H4f 8361 -1.5 -1.8<br />

H1F0 H1 histone family, member 0 3005 -1.0 -1.6<br />

H2AFX H2A histone family, member X 3014 -1.2 -1.5<br />

CCNB1 cyclin B1 891 -1.3 -1.6<br />

CCNB2 cyclin B2 9133 -1.3 -1.6<br />

CCNH cyclin H 902 1.0 -1.7<br />

CCND1 cyclin D1 595 -1.1 1.7<br />

CCNL1 cyclin L1 57018 1.4 2.1<br />

CDK6 cyclin-dependent kinase 6 1021 -1.0 1.6<br />

CDKL2 cyclin-dependent kinase-like 2 (CDC2-related kinase) 8999 1.0 3.2<br />

CDKL5 cyclin-dependent kinase-like 5 6792 1.1 2.6<br />

CDKN1B cyclin-dependent kinase inhibitor 1B (p27, Kip1) 1027 -1.2 -1.6<br />

CDKN2B cyclin-dependent kinase inhibitor 2B (p15, inhibits CDK4) 1030 1.0 -1.9<br />

CDKN3 cyclin-dependent kinase inhibitor 3 1033 -1.3 -1.9<br />

CDKN2C cyclin-dependent kinase inhibitor 2C (p18, inhibits CDK4) 1031 -1.0 1.7<br />

CDCA3 cell division cycle associated 3 83461 -1.4 -2.0<br />

CDCA4 cell division cycle associated 4 55038 -1.0 -1.6<br />

CDCA8 cell division cycle associated 8 55143 -1.5 -2.1<br />

CDC25A cell division cycle 25 homolog A (S. pombe) 993 -1.0 -1.5<br />

CDC14A CDC14 cell division cycle 14 homolog A (S. cerevisiae) 8556 -1.1 2.8<br />

CENPF centromere protein F, 350/400ka (mitosin) 1063 -1.2 -1.9<br />

CEP55 centrosomal protein 55kDa 55165 -1.2 -1.5<br />

CEP97 centrosomal protein 97kDa 79598 -1.1 -1.9<br />

CEP120 centrosomal protein 120kDa 153241 -1.0 -1.6<br />

CEP250 centrosomal protein 250kDa 11190 -1.1 -1.6<br />

CEP170 centrosomal protein 170kDa 9859 -1.1 1.8<br />

CKS2 CDC28 protein kinase regulatory subunit 2 1164 -1.2 -1.7<br />

GAS2L3 growth arrest-specific 2 like 3 283431 -1.4 -2.1<br />

PRC1 protein regulator <strong>of</strong> cytokinesis 1 9055 -1.2 -1.6<br />

SKP2 S-phase kinase-associated protein 2 (p45) 6502 -1.2 -1.5<br />

BTG1 B-cell translocation gene 1, anti-proliferative 694 1.4 1.8<br />

CHEK2 CHK2 checkpoint homolog (S. pombe) 11200 -1.1 3.2<br />

DMTF1 cyclin D binding myb-like transcription factor 1 9988 1.1 1.9<br />

a Fold change values were calculated for the comparison <strong>of</strong> infected GFP + S9 cell with the baseline <strong>of</strong> medium control samples.<br />

Differential regulation is indicated in bold.<br />

124


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

Insights into infection-dependent differential gene expression <strong>by</strong> selected examples from the<br />

6.5 h time point<br />

Infection <strong>of</strong> S9 cells with S. aureus RN1HG and internalization <strong>of</strong> the <strong>pathogen</strong> affected<br />

metabolism related genes 6.5 h after start <strong>of</strong> infection.<br />

Interestingly, the second most strongly induced gene was indoleamine 2,3-dioxygenase 1<br />

(IDO1, fold change 23.5), a key molecule in immunomodulation, microbial growth control, and<br />

<strong>pathogen</strong> immune escape (Zelante et al. 2009). The enzyme catalyzes the reaction <strong>of</strong> tryptophan<br />

to formylkynurenine. In relation to IDO1, the increased expression <strong>of</strong> WARS (4.4) whose gene<br />

product is responsible for tryptophan tRNA charging attracted attention.<br />

Lipid metabolism related genes regulated 6.5 h after start <strong>of</strong> infection in S9 cells featured a<br />

decreased expression in infected samples leading to the assumption <strong>of</strong> a reduced de novo<br />

lipogenesis (Fig. R.4.10).<br />

This concerned different types <strong>of</strong> lipids. In the mevalonate pathway and the following<br />

cholesterol biosynthesis steps, five genes were repressed: mevalonate kinase (MVK; −2.2),<br />

farnesyl-diphosphate farnesyltransferase 1 (FDFT1; −1.7), sterol-C4-methyl oxidase-like (SC4MOL;<br />

−2.0), squalene epoxidase (SQLE; −1.8) and NAD(P) dependent steroid dehydrogenase-like<br />

(NSDHL; −1.7). The induced genes <strong>of</strong> cholesterol 25-hydroxylase (CH25H; 6.0) and oxysterol<br />

binding protein-like 10 (OSBPL10; 1.8) are involved in negative feedback on cholesterol<br />

biosysnthesis. The repression <strong>of</strong> fatty acid synthase (FASN; −1.7) is linked to two genes indirectly<br />

involved in lipid metabolism. Deiodinase type II (DIO2; −4.1) catalyzes the deiodization <strong>of</strong><br />

thyroxine (T 4 ) to the main biologically active form triiodothyronine (T 3 ), which is known to induce<br />

lipogenesis. Although there was probably no hormone T 4 to be converted in the in vitro situation<br />

used for this study, the repression <strong>of</strong> DIO2 might reflect a reaction that makes sense in vivo for<br />

bronchial epithelial cells leading to less local activation <strong>of</strong> lipogenesis. In contrast, glucagon (GCG)<br />

is known to suppress lipogenesis (Hillgartner et al. 1995). This hormone was induced (3.6) in<br />

infected S9 cells. Finally, the two genes for mitochondrial glycerol-3-phosphate acyltransferase<br />

(GPAM; −1.6) and 1-acylglycerol-3-phosphate O-acyltransferase 5 (AGPAT5; −1.5), which are<br />

involved in triacylglycerol biosynthesis, were repressed in infected samples.<br />

Fig. R.4.10:<br />

Repression <strong>of</strong> enzyme genes related to<br />

lipid metabolism.<br />

Lipid metabolism related genes were<br />

chosen with the help <strong>of</strong> omics-viewer(s)<br />

<strong>of</strong> BIOCYC (SRI International, CA, USA,<br />

http://biocyc.org) and manually added<br />

to and arranged in the Ingenuity<br />

Pathway Analysis path designer tool<br />

(IPA, www.ingenuity.com). Green color<br />

indicates repression <strong>of</strong> gene expression<br />

in infected samples at the 6.5 h time<br />

point, red color induction <strong>of</strong> gene<br />

expression, and more intense color<br />

shows a higher absolute fold change.<br />

Enzymes mevalonate kinase (MVK),<br />

farnesyl-diphosphate farnesyltransferase<br />

1 (FDFT1) in mevalonate<br />

pathway, and sterol-C4-methyl oxidaselike<br />

(SC4MOL), SQLE (squalene<br />

epoxidase), and NAD(P) dependent steroid dehydrogenase-like (NSDHL) in the following steps <strong>of</strong> cholesterol biosynthesis were<br />

repressed. CH25H (cholesterol 25-hydroxylase) and OSBPL10 (oxysterol binding protein-like 10), which are involved in negative<br />

feedback on cholesterol biosynthesis, were induced. Fatty acid synthase (FASN), the enzyme <strong>of</strong> fatty acid biosynthesis, was repressed,<br />

and also deiodinase type II (DIO2), an enzyme catalyzing the deiodization <strong>of</strong> thyroxine (T 4) to the main biologically active form<br />

triiodothyronine (T 3). T 3 induces (A) de novo lipogenesis, while glucagon (GCG) suppresses (I) it. The enzymes mitochondrial glycerol-3-<br />

phosphate acyltransferase (GPAM) and 1-acylglycerol-3-phosphate O-acyltransferase 5 (AGPAT5) are involved in triacylglycerol<br />

biosynthesis. Their expression was reduced 6.5 h after start <strong>of</strong> infection.<br />

125


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

Enzymes catalyzing synthesis or degradation reactions <strong>of</strong> lipids are in certain cases also<br />

related to signal transduction and regulatory processes in the cell. In this context, a cumulation <strong>of</strong><br />

increased expression values in infected samples 6.5 h after start <strong>of</strong> infection was noticeable<br />

(Fig. R.4.11).<br />

Phosphatidylinositol-4-kinase and phosphoinositide-3-kinase subunits (PI4K2B, 2.1; PIK3CA,<br />

1.7; PIK3R1, 1.7; PIK3R3, 2.4) belong to the group <strong>of</strong> genes coding for enzymes involved in lipid<br />

messenger reactions. These enzymes catalyze reactions ending with the production <strong>of</strong><br />

phosphatidylinositol-trisphosphate (PIP 3 ), a messenger involved in activation <strong>of</strong> protein kinase<br />

AKT/PKB. Phospholipases C (PLCB4, 2.9; PLCG2, 3.9; PLCH1, 2.0) generate the messengers<br />

inositol-trisphosphate (IP 3 ) and diacylglycerol (DAG). Cytosolic phospholipase A2 gamma<br />

(PLA2G4C, 2.7) releases among others arachidonic acid, which can serve as messenger with the<br />

function <strong>of</strong> directly activating ion channels and protein kinase C (Ordway et al. 1991, Khan WA et<br />

al. 1995, Bonventre 1992) or can be processed to eicosanoid mediators like prostaglandins<br />

(Laye/Gill 2003). Interestingly, the acyl-CoA synthetase (ACSL3, 1.7) with preference for<br />

arachidonate, myristate and eicosapentaenoate was increased, too. This might indicate an<br />

inactivation process for the messenger arachidonic acid. Furthermore, PRKD1 (2.0) and PRKD2<br />

(2.5), coding for protein kinase D alias protein kinase C is<strong>of</strong>orm µ, were induced, too (not shown).<br />

Three more genes showed an increase <strong>of</strong> gene expression 6.5 h after start <strong>of</strong> infection: Serine<br />

palmitoyltransferase subunit (SPTLC2, 1.8) and ketodihydrosphingosine reductase (KDSR, 1.6),<br />

which are part <strong>of</strong> ceramide biosynthesis, and alkaline ceramidase (ACER2, 4.1), an enzyme <strong>of</strong><br />

ceramide degradation to sphingosine. Noticeably, both final products ceramide and sphingosine<br />

are associated with apoptosis.<br />

3-phosphoinositide<br />

biosynthesis<br />

phospholipases<br />

activation <strong>of</strong><br />

fatty acid <strong>by</strong> CoA<br />

ceramide<br />

biosynthesis<br />

sphingosine<br />

metabolism<br />

preference for<br />

arachidonate,<br />

myristate,<br />

eicosapentaenoate<br />

palmityl-CoA<br />

ceramide<br />

PIP 3<br />

IP 3 + DAG<br />

arachidonic<br />

acid<br />

arachidonyl-<br />

CoA<br />

ceramide<br />

sphingosine<br />

activation <strong>of</strong><br />

AKT/PKB<br />

activation <strong>of</strong><br />

calcium-channels<br />

and PKC<br />

activity regulation<br />

<strong>of</strong> ion channels<br />

and PKC<br />

inactivation <strong>of</strong><br />

arachidonic acid<br />

apoptosis<br />

Fig. R.4.11: Induction <strong>of</strong> enzyme genes related to lipid messenger production.<br />

Lipid metabolism related genes were chosen with the help <strong>of</strong> omics-viewer(s) <strong>of</strong> BIOCYC (SRI International, CA, USA, http://biocyc.org)<br />

and manually added to and arranged in the Ingenuity Pathway Analysis path designer tool (IPA, www.ingenuity.com). Red color<br />

indicates induction <strong>of</strong> gene expression in infected samples at the 6.5 h time point, and more intense color shows a higher absolute<br />

fold change. Genes coding for enzymes, which are involved in lipid messenger reactions, were induced: phosphatidylinositol 4-kinase<br />

type 2 beta (PI4K2B), phosphoinositide-3-kinase, catalytic, alpha polypeptide (PIK3CA), phosphoinositide-3-kinase, regulatory subunit<br />

1/alpha (PIK3R1), and phosphoinositide-3-kinase, regulatory subunit 3/gamma (PIK3R3). These enzymes catalyze reactions ending<br />

with the production <strong>of</strong> phosphatidylinositol-trisphosphate (PIP 3), a messenger involved in activation <strong>of</strong> protein kinase AKT/PKB.<br />

Phospholipase C, beta 4 (PLCB4), phosphatidylinositol-specific phospholipase C, gamma 2 (PLCG2), and phospholipase C, eta 1 (PLCH1)<br />

generate the messengers inositol-trisphosphate (IP 3) and diacylglycerol (DAG), cytosolic, calcium-independent phospholipase A2,<br />

group IV C (PLA2G4C) generates the messenger arachidonic acid. Acyl-CoA synthetase long-chain family member 3 (ACSL3) was<br />

increased, too. This might indicate in inactivation process for the messenger arachidonic acid.<br />

Serine palmitoyltransferase, long chain base subunit 2 (SPTLC2) and 3-ketodihydrosphingosine reductase (KDSR), which are part <strong>of</strong><br />

ceramide biosynthesis, and alkaline ceramidase 2 (ACER2), an enzyme <strong>of</strong> ceramide degradation to sphingosine, showed a higher<br />

expression in infected samples than in controls at the 6.5 h time point.<br />

126


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

Differential expression <strong>of</strong> apoptosis associated genes has already been observed in the data<br />

evaluation with Ingenuity Pathway Analysis tools (see pathway “Death Receptor Signaling”,<br />

page 120). Further apoptosis related genes were found when manually filtering the list <strong>of</strong><br />

differentially expressed genes at the 6.5 h time point (Table R.4.7). Genes for apoptosis inducing<br />

proteins B-cell CLL/lymphoma 10 (BCL10, 1.7) and apoptosis facilitator BCL2-like 11 (BCL2L11,<br />

2.0) were increased in infected samples. The proteins belong to the group <strong>of</strong> BH3-only proteins<br />

and achieve their apoptotic effect <strong>by</strong> binding to proteins <strong>of</strong> the Bcl-2 family. The mechanism is<br />

discussed to take effect <strong>by</strong> activating pro-apoptotic members or inactivating pro-survival<br />

members in a de-repression mode (Bouillet/Strasser 2002, Adams 2003, Fletcher/Huang 2006).<br />

Also apolipoprotein L1 (APOL1) is a BH3-only protein whose accumulation leads to autophagy<br />

(Wan et al. 2008, Zhaorigetu et al. 2008) and is postulated to be linked to apoptosis<br />

(Vanhollebeke/Pays 2006). Its expression was increased <strong>by</strong> a factor <strong>of</strong> 3.4 in infected samples at<br />

the 6.5 h time point. Interestingly, four other apolipoprotein L genes, which do not possess a<br />

secretion signal peptide like APOL1 and therefore are thought to be localized intracellularly<br />

(Duchateau et al. 1997, Page et al. 2001), showed increased expression: APOL2 (4.6), APOL3 (3.7),<br />

APOL4 (1.9), and APOL6 (4.2). Expression <strong>of</strong> APOL5 was absent (p > 0.01 on intensity pr<strong>of</strong>ile level<br />

in Rosetta Resolver analysis) in all 16 infected and control S9 cell samples <strong>of</strong> this study. These<br />

other APOL proteins also contain BH3-domains and are supposed to be associated with<br />

programmed cell death and immune response (Liu Z et al. 2005). Most interestingly, APOL1 is the<br />

target protein <strong>of</strong> the antagonistic Trypanosoma brucei rhodesiense protein SRA which helps the<br />

<strong>pathogen</strong> to evade the immune response. Further pro-apoptotic genes were induced in infected<br />

S9 cells like BH3-like motif containing, cell death inducer (BLID; 2.3), BCL2-antagonist/killer 1<br />

(BAK1; 1.9), caspase 4, apoptosis-related cysteine peptidase (CASP4; 2.8), but also an antiapoptotic<br />

gene, BCL2-associated athanogene (BAG1; 1.8; Table R.4.7).<br />

Table R.4.7: Overview on selected differentially expressed apoptosis related genes in S9 cells 6.5 h after start <strong>of</strong> infection with<br />

S. aureus RN1HG GFP.<br />

Rosetta Resolver annotation<br />

fold change a<br />

gene<br />

name<br />

description<br />

Entrez<br />

Gene ID<br />

BCL10 B-cell CLL/lymphoma 10 8915 CLAP, mE10, CIPER, c-E10, CARMEN 1.7<br />

BCL2L11 BCL2-like 11 (apoptosis facilitator) 10018<br />

BAM, BIM, BOD, BimL, BimEL, BIMbeta6,<br />

BIM-beta7, BIM-alpha6<br />

2.0<br />

APOL1 apolipoprotein L, 1 8542 APOL, APO-L, APOL-I 3.4<br />

APOL2 apolipoprotein L, 2 23780 APOL3, APOL-II 4.5<br />

APOL3 apolipoprotein L, 3 80833 CG12-1, APOLIII 3.7<br />

APOL4 apolipoprotein L, 4 80832 APOLIV,APOL-IV 1.9<br />

APOL6 apolipoprotein L, 6 80830 APOLVI, APOL-VI 4.2<br />

BLID BH3-like motif containing, cell death inducer 414899 BRCC2 2.3<br />

BAK1 BCL2-antagonist/killer 1 578 CDN1, BCL2L7, BAK-LIKE 1.9<br />

CASP4 caspase 4, apoptosis-related cysteine peptidase 837 TX, ICH-2, Mih1/TX, ICEREL-II 2.8<br />

BAG1 BCL2-associated athanogene 573 RAP46 1.8<br />

a Fold change values were calculated for the comparison <strong>of</strong> infected GFP + S9 cell with the baseline <strong>of</strong> medium control samples.<br />

alias<br />

6.5 h<br />

Manual filtering revealed induced cytokines and cytokine receptors 6.5 h after start <strong>of</strong><br />

infection, some <strong>of</strong> which have already been mentioned in the context <strong>of</strong> IFN signaling or pattern<br />

recognition receptors (Table R.4.8). Only IFNB1 and IL6 were already induced at the 2.5 h time<br />

point. In total, the induced genes revealed a pro-inflammatory response: the acute phase<br />

response and fever mediator IL-6, inducers <strong>of</strong> MHC-I molecules and antiviral/antibacterial<br />

mechanisms (IFNB1, IL28B), chemoattractants for monocytes, T cells, denditic cells and<br />

granulocytes (CCL2, CCL5, CXCL10, CXCL11, CXCL16), B and T cell activating cytokines (IL-7,<br />

TNFSF13B), macrophage growth factor (CSF1), inducer <strong>of</strong> other cytokines and immune cell<br />

127


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

differentiation (IL12A), epithelial receptor for the inflammatory cytokine IL-22 (IL22RA1),<br />

scavenger receptor for cytokines (CCRL1), immunomodulatory cytokines (IL28A, IL28B, IL29), and<br />

apoptosis inducers (TNFSF10, TNFSF15), but also anti-apoptotic IL-15 and its receptor IL15RA.<br />

Table R.4.8: Overview on differentially expressed cytokines and receptors in S9 cells 6.5 h after start <strong>of</strong> infection with S. aureus RN1HG<br />

GFP.<br />

Rosetta Resolver annotation<br />

fold change a<br />

gene<br />

name<br />

description<br />

Entrez<br />

Gene ID<br />

CCL2 chemokine (C-C motif) ligand 2 6347 MCP1 2.8<br />

CCL5 chemokine (C-C motif) ligand 5 6352 RANTES 4.0<br />

CCRL1 chemokine (C-C motif) receptor-like 1 51554 CCR11 2.3<br />

CSF1 colony stimulating factor 1 (macrophage) 1435 MCSF 3.1<br />

CXCL10 chemokine (C-X-C motif) ligand 10 3627 IP-10, INP10, IFI10 28.1<br />

CXCL11 chemokine (C-X-C motif) ligand 11 6373 IP9, I-TAC 7.0<br />

CXCL16 chemokine (C-X-C motif) ligand 16 58191 SR-PSOX 1.8<br />

CXCR4 chemokine (C-X-C motif) receptor 4 7852 CD184, SDF1R, LESTR 1.6<br />

IFNB1 interferon, beta 1, fibroblast 3456 3.6<br />

IL6 interleukin 6 (interferon, beta 2) 3569 HGF,HSF,BSF2,IFNB2 11.9<br />

IL7 interleukin 7 3574 2.0<br />

IL12A interleukin 12A 3592 P35,CLMF,NKSF1 4.0<br />

IL15 interleukin 15 3600 3.4<br />

IL15RA interleukin 15 receptor, alpha 3601 2.1<br />

IL22RA1 interleukin 22 receptor, alpha 1 58985 1.8<br />

IL28A interleukin 28A (interferon, lambda 2) 282616 IFNL2 5.1<br />

IL28B interleukin 28B (interferon, lambda 3) 282617 IFNL3 3.0<br />

IL29 interleukin 29 (interferon, lambda 1) 282618 IFNL1 5.7<br />

TNFSF10 tumor necrosis factor (ligand) superfamily, member 10 8743 TRAIL, TL2, APO2L, CD253 18.5<br />

TNFSF13B tumor necrosis factor (ligand) superfamily, member 13b 10673 DTL, BAFF, BLYS, CD257 2.2<br />

TNFSF15 tumor necrosis factor (ligand) superfamily, member 15 9966 TL1, VEGI 4.6<br />

a Fold change values were calculated for the comparison <strong>of</strong> infected GFP + S9 cell with the baseline <strong>of</strong> medium control samples.<br />

alias<br />

6.5 h<br />

Further interesting induced genes included myxovirus (influenza virus) resistance 1 and 2<br />

(interferon-inducible protein p78, MX1; 2.2; MX2; 1.9). These genes code for interferon-induced,<br />

antiviral proteins forming a ring-shaped oligomeric structure, which recognizes and sequesters<br />

viral nucleocapsid proteins (Gao et al. 2010). MX1 and MX2 belong to the same family <strong>of</strong><br />

interferon-inducible GTPases as the four guanylate binding proteins 1, 3, 4, 5 (interferoninducible,<br />

67kDa, GBP1; 2.6; GBP2; 2.9; GBP3; 14.9; GBP4; 11.1), which were induced in S9 cells<br />

6.5 h after start <strong>of</strong> infection with S. aureus RN1HG GFP. These GBPs are involved in immune<br />

defense, probably oligomerize, act antivirally and inhibit cellular proliferation (MacMicking 2004).<br />

The fifth member <strong>of</strong> the p65 GBP family, GBP6, was not expressed on all 16 arrays in this study<br />

(Table R.4.9).<br />

The endothelial protein C receptor (EPCR; PROCR; 1.6) was also induced in infected S9 cells.<br />

Protein C activation <strong>by</strong> thrombin/thrombomodulin is enhanced when it is bound to EPRC, and<br />

subsequently, activated protein C can act in an anticoagulant and anti-inflammatory manner. In<br />

inflammation, where coagulation is facilitated <strong>by</strong> different mechanisms, e. g. via C-reactive<br />

protein CRP, the induction <strong>of</strong> PROCR might indicate a counterregulatory process (Esmon 2006).<br />

Interestingly, the receptor shedding from the cellular surface is mediated <strong>by</strong> tumor necrosis<br />

factor-alpha converting enzyme / TACE, alias ADAM metallopeptidase domain 17 / ADAM17 (Qu et<br />

al. 2006), which was also induced in the infected S9 cells (ADAM17; 1.8; Table R.4.9).<br />

Further gene expression changes in infected S9 cells could be assigned to a related<br />

physiological aspect: Tissue factor pathway inhibitor 2 (TFPI2; 2.2) belongs like activated<br />

protein C to the natural anticoagulants (Esmon 2006).<br />

128


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

Plasminogen activator urokinase (PLAU, −2.0) activates plasmin and therefore leads to anticoagulant<br />

activity. But its proteolytic activity also leads to degradation <strong>of</strong> extracellular matrix<br />

(Dass et al. 2008). The expression <strong>of</strong> PLAU was decreased in infected S9 cells. Contrarily,<br />

plasminogen activator urokinase receptor (PLAUR; 4.0), receptor for PLAU, was induced. The<br />

receptor promotes and localizes extracellular protease activity and plasmin formation <strong>by</strong> PLAU,<br />

but also participates in internalization <strong>of</strong> complexes between PLAU and its inhibitors (Cubellis et<br />

al. 1990, Dass et al. 2008, Blasi/Sidenius 2009). Plasminogen activator inhibitor-1 (PAI-1) is the<br />

main inhibitor <strong>of</strong> PLAU, and it was induced in infected S9 cells (SERPINE1; 3.3; Table R.4.9)<br />

In infected S9 cells, induction was observed for mitochondrial superoxide dismutase 2 (SOD2;<br />

2.6), a gene known to be inducible in response to oxidative stress and <strong>by</strong> inflammatory cytokines<br />

like IL-6 (Dougall/Nick 1991, Miao et al. 2009). Furthermore, thioltransferase glutaredoxin (GLRX;<br />

1.9) and thioredoxin interacting protein (TXNIP; 2.2), which are involved in anti-oxidant defense,<br />

were induced (Table R.4.9).<br />

Four components <strong>of</strong> the complement system were induced in infected S9 cells at the 6.5 h<br />

time point: complement component 1, r subcomponent-like and s subcomponent (C1RL; 1.7;<br />

C1S; 1.7) <strong>of</strong> the classical pathway, complement factor B (CFB; 3.1) <strong>of</strong> the alternative pathway and<br />

complement component 3a receptor 1 (C3AR1; 3.6; Table R.4.9).<br />

Lysosomal enzyme genes <strong>of</strong> cathepsin S (CTSS; 2.6) and legumain (LGMN; 2.3) and<br />

additionally, lysosomal membrane protein genes LAMP3 (5.7) and LAPTM5 (1.9) were induced.<br />

Table R.4.9: Overview on selected differentially expressed immune defense related genes in S9 cells 6.5 h after start <strong>of</strong> infection with<br />

S. aureus RN1HG GFP.<br />

Rosetta Resolver annotation<br />

fold change a<br />

gene name<br />

description<br />

Entrez<br />

Gene ID<br />

MX1 myxovirus (influenza virus) resistance 1, interferoninducible<br />

4599 MxA, IFI78 2.2<br />

protein p78 (mouse)<br />

MX2 myxovirus (influenza virus) resistance 2 (mouse) 4600 MXB 1.9<br />

GBP1 guanylate binding protein 1, interferon-inducible, 67kDa 2633 2.6<br />

GBP3 guanylate binding protein 3 2635 2.9<br />

GBP4 guanylate binding protein 4 115361 Mpa2 14.9<br />

GBP5 guanylate binding protein 5 115362 11.1<br />

PROCR protein C receptor, endothelial (EPCR) 10544 EPCR, CCD41, CD201 1.6<br />

ADAM17 ADAM metallopeptidase domain 17 6868 CSVP, TACE, CD156B 1.8<br />

TFPI2 tissue factor pathway inhibitor 2 7980 PP5, REF1 2.2<br />

PLAU plasminogen activator, urokinase 5328 ATF, UPA, URK, u-PA -2.0<br />

PLAUR plasminogen activator, urokinase receptor 5329 CD87, UPAR, URKR 4.0<br />

SERPINE1 serpin peptidase inhibitor, clade E (nexin, plasminogen 5054 PAI, PAI-1, PLANH1 3.3<br />

activator inhibitor type 1), member 1<br />

SOD2 superoxide dismutase 2, mitochondrial 6648 IPO-B, Mn-SOD 2.6<br />

GLRX glutaredoxin (thioltransferase) 2745 GRX, GRX1 1.9<br />

TXNIP thioredoxin interacting protein 10628 TXNIP, THIF, VDUP1 2.2<br />

C1RL complement component 1, r subcomponent-like 51279 1.7<br />

C1S complement component 1, s subcomponent 716 1.7<br />

CFB complement factor B 629 3.1<br />

C3AR1 complement component 3a receptor 1 719 3.6<br />

CTSS cathepsin S 1520 2.6<br />

LGMN legumain 5641 AEP, PRSC1 2.3<br />

LAMP3 lysosomal-associated membrane protein 3 27074 CD208, TSC403 5.7<br />

LAPTM5 lysosomal multispanning membrane protein 5 7805 CLAST6 1.9<br />

a Fold change values were calculated for the comparison <strong>of</strong> infected GFP + S9 cell with the baseline <strong>of</strong> medium control samples.<br />

alias<br />

6.5 h<br />

Several adhesins were induced in S. aureus RN1HG GFP infected S9 cells, like integrins alpha<br />

and beta (ITGA2; 2.9; ITGA5; 2.0; ITGA11; 1.6; ITGB8; 2.6) and their regulator cytohesin 1 (CYTH1;<br />

2.1), protocadherins (PCDH7; 4.1; PCDH17; 2.1) and cadherin CDH8 (1.6). Another cadherin was<br />

129


Maren Depke<br />

Results<br />

Host Cell Gene Expression Pattern in an in vitro Infection Model<br />

repressed (CDH10, −1.9). Increased expression was also observed for adhesion molecules ALCAM<br />

(1.8) and CEACAM1 (9.9). ALCAM was one <strong>of</strong> the genes whose protein abundance change was in<br />

accordance with the differential expression (Table R.4.10, Table R.4.3).<br />

Table R.4.10: Overview on selected differentially expressed cell adhesion related genes in S9 cells 6.5 h after start <strong>of</strong> infection with<br />

S. aureus RN1HG GFP.<br />

gene<br />

name<br />

description<br />

Rosetta Resolver annotation<br />

Entrez<br />

Gene<br />

ID<br />

alias<br />

fold change a<br />

ITGA2 integrin, alpha 2 (CD49B, alpha 2 subunit <strong>of</strong> VLA-2 receptor) 3673 BR, GPIa, CD49B,VLA-2 2.9<br />

ITGA5 integrin, alpha 5 (fibronectin receptor, alpha polypeptide) 3678 FNRA, CD49e,VLA5A 2.0<br />

ITGA11 integrin, alpha 11 22801 1.6<br />

ITGB8 integrin, beta 8 3696 ITGB8 2.6<br />

ITGB1BP2 integrin beta 1 binding protein (melusin) 2 26548 CHORDC3, ITGB1BP 1.8<br />

CYTH1 cytohesin 1 9267 B2-1,SEC7,PSCD1 2.1<br />

PCDH7 protocadherin 7 5099 BHPCDH 4.1<br />

PCDH17 protocadherin 17 27253 PCH68, PCDH68 2.1<br />

CDH8 cadherin 8, type 2 1006 1.6<br />

CDH10 cadherin 10, type 2 (T2-cadherin) 1008 -1.9<br />

ALCAM activated leukocyte cell adhesion molecule 214 MEMD, CD166 1.8<br />

CEACAM1<br />

carcinoembryonic antigen-related cell adhesion molecule 1<br />

(biliary glycoprotein)<br />

634 BGP, BGP1, BGPI 9.9<br />

a Fold change values were calculated for the comparison <strong>of</strong> infected GFP + S9 cell with the baseline <strong>of</strong> medium control samples.<br />

6.5 h<br />

130


Maren Depke<br />

Results<br />

PATHOGEN GENE EXPRESSION PROFILING<br />

Growth Media Comparison Study<br />

Reproducibility <strong>of</strong> replicates and clustering <strong>of</strong> experimental condition groups<br />

Cultivation <strong>of</strong> S. aureus RN1HG in pMEM medium was performed in triplicate. In the tiling<br />

array analysis <strong>of</strong> the sample points exponential growth, stationary phase t 2 , and stationary<br />

phase t 4 , the replicates <strong>of</strong> each group were arranged together in a cluster. As expected, for the<br />

stationary t 4 time point, which is defined as 4 h after entry into stationary phase, a higher<br />

similarity to the t 2 samples, which were harvested 2 h earlier, than to the exponential growth<br />

samples was detected. This similarity was especially visible for the t 4 biological replicates 2 and 1,<br />

whereas the replicate 3 held more distance to the other samples (Fig. R.5.1).<br />

exponentialgrowth<br />

stationary phase t 2<br />

stationary phase t 4<br />

biological replicate 1<br />

biological replicate 2<br />

biological replicate 3<br />

Fig. R.5.1: Hierarchical clustering <strong>of</strong> 9 tiling array data sets from growth in pMEM medium.<br />

The following clustering algorithms were applied on z-score-transformed data: Agglomerative clustering with average linkage using<br />

cosine correlation as similarity measure. All sequences were included in the cluster analysis.<br />

Three major classes according to the sample points during growth were discernible: 1) exponential growth, 2) stationary phase t 2, and<br />

3) stationary phase t 4. As expected for the stationary t 4 time point, a higher similarity to the t 2 samples than to the exponential growth<br />

samples was detected. This similarity was especially visible for the t 4 biological replicates 2 and 1, whereas the replicate 3 held more<br />

distance to the other samples.<br />

Comparison <strong>of</strong> experimental condition groups and assessment <strong>of</strong> differentially regulated genes<br />

The consumption <strong>of</strong> medium components and the reduction <strong>of</strong> growth rate after beginning <strong>of</strong><br />

stationary phase was accompanied <strong>by</strong> a substantial change <strong>of</strong> gene expression. This change was<br />

visualized <strong>by</strong> the strong scattering when comparing stationary phase t 2 and stationary phase t 4<br />

samples to the exponential growth in scatter plots (Fig. R.5.2).<br />

The comparison <strong>of</strong> stationary phase samples with exponential growth samples <strong>by</strong> statistical<br />

testing resulted in 2243 and 1399 sequences, which exhibited a significant difference to<br />

exponential growth at the t 2 and the t 4 time point, respectively. Of these, 1423 sequences<br />

exceeded an absolute fold change <strong>of</strong> 2 in the t 2 samples, and 1086 sequences passed the cut<strong>of</strong>f in<br />

the 2 h later samples <strong>of</strong> the t 4 time point (Table R.5.1).<br />

131


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

<br />

stationary phase t 2<br />

<br />

stationary phase t 4<br />

exponential growth<br />

exponential growth<br />

Fig. R.5.2: Scatter plots comparing mean signal intensities <strong>of</strong> treatment groups.<br />

The signals <strong>of</strong> the three groups <strong>of</strong> “stationary phase t 2”, “stationary phase t 4”, and “exponential growth phase” were plotted after<br />

combining the three biological replicates. All sequences available on the array are shown.<br />

Table R.5.1: Results <strong>of</strong> group comparisons with statistical testing <strong>of</strong> stationary phase and exponential growth staphylococcal array<br />

data sets.<br />

number <strong>of</strong> sequences<br />

group comparison a<br />

significant with p* < 0.05 in textbook oneway<br />

ANOVA with Benjamini-Hochberg False<br />

Discovery Rate multiple testing correction<br />

absolute fold change equal<br />

to or greater than 2 AND<br />

significant with p* < 0.05<br />

stationary phase t 2 vs. exponential growth phase 2243 1423<br />

stationary phase t 4 vs. exponential growth phase 1399 1086<br />

a The 080604_SA_JH_Tiling array contains 3882 sequences <strong>of</strong> which 2825 are known transcripts. The remaining 1057 sequences were<br />

identified as new transcripts in the tiling array analysis. All sequences were included in statistical testing.<br />

Comparison <strong>of</strong> differentially regulated genes in t 2 and t 4 <strong>of</strong> stationary phase in pMEM<br />

The gene expression signatures <strong>of</strong> t 2 and t 4 stationary phase S. aureus RN1HG, which resulted<br />

from statistical comparison with the baseline <strong>of</strong> exponential growth including multiple testing<br />

correction and minimal absolute fold change cut<strong>of</strong>f 2, were compared in a Venn diagram<br />

(Fig. R.5.3 A). For 940 sequences, differential expression was observed in both time points <strong>of</strong><br />

stationary phase, while 483 and 146 sequences were specific for the t 2 and t 4 stationary phase<br />

signature, respectively.<br />

This corresponds to a fraction <strong>of</strong> 66 % <strong>of</strong> sequences regulated at t 2 , which were also<br />

differentially expressed at t 4 . Vice versa, 87 % <strong>of</strong> differentially expressed sequences at the t 4 time<br />

point were also found to be regulated at t 2 . When examining the induced and repressed<br />

sequences separately, it was first obvious that sequences regulated at both time points always<br />

possessed the same regulation direction in these two time points (Fig. R.5.3 B). Further, a slightly<br />

higher number <strong>of</strong> repressed sequences was visible in the differentially expressed sequences, but<br />

induction accounted for almost the same number <strong>of</strong> sequences as repression.<br />

In the t 2 samples, 696 sequences were induced (225 specifically for that time point), 727 were<br />

repressed (258 specifically for that time point), while in the t 4 samples 540 exhibited induction<br />

(69 specifically for that time point) and 546 exhibited repression (77 specifically for that time<br />

point). Therefore, the 940 sequences differentially expressed in both time points consisted <strong>of</strong> 471<br />

induced and 469 repressed sequences.<br />

132


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

A<br />

B<br />

483 940 146<br />

repressed<br />

repressed<br />

t 2<br />

t 4<br />

induced<br />

258 77<br />

induced<br />

225<br />

469<br />

69<br />

sequences differentially expressed<br />

between stationary growth phase t 2<br />

and exponential growth phase <strong>of</strong><br />

S. aureus RN1HG cultivated in pMEM<br />

sequences differentially expressed<br />

between stationary growth phase t 4<br />

and exponential growth phase <strong>of</strong><br />

S. aureus RN1HG cultivated in pMEM<br />

471<br />

Fig. R.5.3: Comparison <strong>of</strong> the t 2 and t 4 stationary phase signatures <strong>of</strong> S. aureus RN1HG in pMEM.<br />

Both stationary phase samples were compared to the baseline <strong>of</strong> exponential growth with statistical testing and multiple testing<br />

correction (p* < 0.05), and a minimal absolute fold change cut<strong>of</strong>f <strong>of</strong> 2 was applied. The comparison <strong>of</strong> differentially expressed<br />

sequences at the t 2 and t 4 time point was performed for all regulated sequences (A) and for induced and repressed sequences<br />

separately (B).<br />

Fractions <strong>of</strong> known genes and newly identified transcribed sequences included on the tiling<br />

array and in the lists <strong>of</strong> differentially expressed sequences derived from the comparisons <strong>of</strong><br />

different growth phase samples<br />

The 080604_SA_JH_Tiling array contains 3882 sequences <strong>of</strong> which 2825, corresponding to<br />

73 %, are known transcripts and associated with a LocusTag number (SAOUHSC_*). The remaining<br />

1057 sequences were identified as new transcripts in the tiling array analysis. These new<br />

transcripts are expected to belong to all types <strong>of</strong> RNAs like mRNA and regulatory RNA, but also<br />

formerly unknown transcribed fragments 5’ and 3’ to known genes might be included.<br />

Furthermore, some artifacts might still be included. All sequences, known and new, were<br />

included in statistical testing.<br />

For the stationary phase-specific signature, 444 and 307 differentially expressed transcripts<br />

belonged to the group <strong>of</strong> new transcripts at the t 2 and t 4 time point, respectively (Table R.5.2).<br />

Table R.5.2: Fractions <strong>of</strong> known annotated and newly detected transcripts in the results <strong>of</strong> group comparisons with statistical testing<br />

<strong>of</strong> different growth phase array data sets.<br />

group comparison a<br />

total number <strong>of</strong><br />

sequences with<br />

absolute fold change<br />

equal to or greater<br />

than 2 AND significant<br />

with p* < 0.05<br />

number <strong>of</strong> known<br />

transcribed sequences<br />

with absolute fold<br />

change equal to or<br />

greater than 2 AND<br />

significant with p* < 0.05<br />

number <strong>of</strong> new<br />

transcribed sequences<br />

with absolute fold<br />

change equal to or<br />

greater than 2 AND<br />

significant with p* < 0.05<br />

stationary phase t 2 vs. exponential growth phase 1423 979 444<br />

stationary phase t 4 vs. exponential growth phase 1086 779 307<br />

a The 080604_SA_JH_Tiling array contains 3882 sequences <strong>of</strong> which 2825 are known transcripts. The remaining 1057 sequences were<br />

identified as new transcripts in the tiling array analysis. All sequences were included in statistical testing.<br />

Physiological aspects <strong>of</strong> stationary phase response<br />

Stationary phase is characterized <strong>by</strong> a strongly reduced growth rate and <strong>by</strong> a reorganization <strong>of</strong><br />

the metabolic processes <strong>of</strong> the bacterial cell. Causative for these alterations is the consumption<br />

<strong>of</strong> nutrients <strong>of</strong> the growth medium and the following starvation for the preferred carbon source.<br />

Such changes have been published for S. aureus COL during growth in TSB in a global gel-based<br />

and gel-free proteome study <strong>by</strong> Kohler et al. in 2005.<br />

133


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

The tiling array approach aimed at the identification <strong>of</strong> the maximal possible number <strong>of</strong><br />

transcriptional units − known and newly identified – and therefore included several different<br />

growth media in a cooperation between several laboratories. On the other hand, the tiling array<br />

analysis could be used to get insights into the stationary phase response in the newly established<br />

pMEM medium on transcriptome level under restriction to the already defined transcriptional<br />

units and annotated genes. For this purpose, differential expression was assumed for genes<br />

regulated in at least one <strong>of</strong> the two analyzed time points <strong>of</strong> stationary phase, t 2 and t 4 .<br />

Using this application <strong>of</strong> the array data set, a very clear picture <strong>of</strong> induced TCA cycle enzyme<br />

genes (citZ, citB, citC, sucA, sucB, sucD, sucC, sdhA, sdhB, sdhC; all induced) was revealed. In<br />

parallel, reduced expression <strong>of</strong> glycolysis enzyme genes (pgi, pfkA, pgk, pgm, pykA; all repressed)<br />

and induced expression <strong>of</strong> gluconeogenetic enzyme genes (pckA, gap2, fbp) was observed. This is<br />

in accordance with published stationary phase response <strong>of</strong> S. aureus (Kohler et al. 2005).<br />

Furthermore, repression <strong>of</strong> amino acid biosynthesis pathways was described for the stationary<br />

phase. In pMEM, tyryptophan (trpB, trpC, trpD, trpE, trpG), histidine (hisB, hisD, hisG), and<br />

aspartate/arginine (argG, argH, SAOUHSC_00150) biosynthesis genes were repressed. But<br />

contrarily, induction <strong>of</strong> lysine (lysC, asd), histidine/glutamate (hutH, hutI, hutU, hutG, rocA), and<br />

citruline/ornithine (argF, arg, arcC) biosynthesis genes was visible in the stationary phase<br />

samples cultivated in pMEM. Other amino acid biosynthesis genes (leu, ilv, dap, thr, and others)<br />

were not significantly differentially expressed. Reduced or even ceased growth diminishes the<br />

need for new protein synthesis. Therefore, associated molecules like ribosomal proteins,<br />

translation elongation factors, and chaperones do not need to be newly synthesized, which was<br />

observed on proteome level for S. aureus COL in TSB (Kohler et al. 2005). This phenomenon is<br />

part <strong>of</strong> the stringent response e. g. to glucose starvation. Here, using pMEM medium, S. aureus<br />

RN1HG and transcriptome analysis, only two ribosomal protein genes, rpsD and rpsT, and prmA, a<br />

ribosomal protein L11 methyltransferase, were repressed. Only translation elongation factor P<br />

(efp) exhibited a fold change <strong>of</strong> less than −2 in the t 4 samples, but this repression was not<br />

significant. Chaperones dnaK, grpE, groEL, and groES exhibited slightly decreased fold change<br />

values in the range <strong>of</strong> −1.4 to −1.9, but the difference was also not significant. Of the Clp<br />

complex, expression <strong>of</strong> subunit X was repressed and <strong>of</strong> subunit L was induced. Finally, repression<br />

<strong>of</strong> tRNA synthetases is known to be part <strong>of</strong> the stringent response in S. aureus (Anderson KL et al.<br />

2006). In stationary phase in pMEM, repression <strong>of</strong> tyrS, leuS, alaS, aspS, hisS, valS, thrS, glyS,<br />

proS, ileS, pheS, pheT, gltX, metS/metG, and serS was detected. The other tRNA synthetases<br />

except cysS showed a trend <strong>of</strong> repression with fold change values almost reaching the cut<strong>of</strong>f (−2).<br />

Virulence factors differentially expressed in stationary vs. exponential growth phase<br />

In stationary growth phase, differential expression <strong>of</strong> virulence-associated genes was<br />

observed. Many <strong>of</strong> these genes were regulated at both analyzed time points <strong>of</strong> stationary phase,<br />

t 2 and t 4 . Surface proteins A and G (spa, sasG) were repressed. Other membrane-bound adhesins<br />

were induced like clfA, fnbA, and isaB. Secreted adhesins and immunomodulatory molecules efb,<br />

chp, and sbi were repressed, whereas ebh and eap were induced. One <strong>of</strong> the two staphylococcal<br />

superoxide dismutases, sodM, was found to be induced. The toxins hla, hlb, hlgC, and hlgB<br />

exhibited an increase in expression. In the group <strong>of</strong> extracellular enzymes, only nuc and htrA<br />

were repressed. Contrarily, for secreted lipases geh and lip and for proteases sspB, splC, splB,<br />

splA, and aur a higher expression was detected in stationary phase than in exponential growth.<br />

The complete cap operon <strong>of</strong> 15 capsular genes (SAOUHSC_00114 to SAOUHSC_00128) was<br />

induced in t 2 samples and still 10 <strong>of</strong> these genes were also induced in t 4 samples. The bi<strong>of</strong>ilm<br />

repressor icaR was induced in t 2 samples. Fittingly, two genes <strong>of</strong> the ica operon, icaB and icaC,<br />

were observed to be repressed in both analyzed stationary phase time points. Finally, differential<br />

expression <strong>of</strong> five staphylococcal accessory regulators was detected: sarX and sarT were<br />

repressed, while sarV, sarR, and sarZ were induced in stationary phase (Table R.5.3).<br />

134


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

Table R.5.3: Expression <strong>of</strong> virulence associated genes in stationary phase in comparison to exponential growth in pMEM.<br />

LocusTag gene annotation<br />

fold change in<br />

comparison to baseline<br />

exponential growth<br />

stationary<br />

phase t 2 a<br />

stationary<br />

phase t 4 a<br />

SAOUHSC_00069 spa protein A -7.4 -8.8<br />

SAOUHSC_02798 sasG surface protein G -4.2 -3.4<br />

SAOUHSC_00812 clfA clumping factor 2.8 2.1<br />

SAOUHSC_02803 fnbA fibronectin-binding protein precursor, putative 3.1 2.7<br />

SAOUHSC_02972 isaB immunodominant staphylococcal antigen B 3.2 3.4<br />

SAOUHSC_01114 efb fibrinogen-binding protein -2.7 1.4<br />

SAOUHSC_02169 chp chemotaxis inhibitory protein -4.3 -2.6<br />

SAOUHSC_02706 sbi immunoglobulin G-binding protein Sbi, putative -3.9 -1.7<br />

SAOUHSC_01447 ebh extracellular matrix-binding protein ebh 4.7 2.9<br />

SAOUHSC_02161 eap MHC class II analog protein 2.5 3.9<br />

SAOUHSC_00093 sodM superoxide dismutase, putative 2.1 1.6<br />

SAOUHSC_01121 hla alpha-hemolysin precursor 15.1 36.8<br />

SAOUHSC_02163 hlb hypothetical protein SAOUHSC_02163 1.9 7.2<br />

SAOUHSC_02709 hlgC leukocidin s subunit precursor, putative 16.2 43.6<br />

SAOUHSC_02710 hlgB leukocidin f subunit precursor 11.5 29.6<br />

SAOUHSC_00818 nuc thermonuclease precursor -2.4 -1.4<br />

SAOUHSC_00958 htrA serine protease HtrA, putative -5.7 -6.3<br />

SAOUHSC_00300 geh lipase precursor 10.4 11.1<br />

SAOUHSC_03006 lip lipase 70.4 97.8<br />

SAOUHSC_00987 sspB cysteine protease precursor, putative 2.1 1.8<br />

SAOUHSC_01939 splC serine protease SplC 2.8 5.2<br />

SAOUHSC_01941 splB serine protease SplB 3.1 6.3<br />

SAOUHSC_01942 splA serine protease SplA 2.9 6.0<br />

SAOUHSC_02971 aur aureolysin, putative 2.9 5.0<br />

SAOUHSC_00114 capA capsular polysaccharide biosynthesis protein, putative 7.0 6.4<br />

SAOUHSC_00115 capB capsular polysaccharide synthesis enzyme Cap5B 6.5 5.5<br />

SAOUHSC_00116 capC capsular polysaccharide synthesis enzyme Cap8C 6.0 5.2<br />

SAOUHSC_00117 capD capsular polysaccharide biosynthesis protein Cap5D, putative 5.7 4.9<br />

SAOUHSC_00118 capE capsular polysaccharide biosynthesis protein Cap5E, putative 5.3 4.2<br />

SAOUHSC_00119 capF capsular polysaccharide synthesis enzyme Cap8F 4.8 3.8<br />

SAOUHSC_00120 capG UDP-N-acetylglucosamine 2-epimerase 4.1 3.4<br />

SAOUHSC_00121 capH<br />

capsular polysaccharide synthesis enzyme O-acetyl transferase Cap5H,<br />

putative<br />

3.9 3.2<br />

SAOUHSC_00122 capI capsular polysaccharide biosynthesis protein Cap5I, putative 3.7 2.9<br />

SAOUHSC_00123 capJ capsular polysaccharide biosynthesis protein Cap5J, putative 3.2 2.3<br />

SAOUHSC_00124 capK capsular polysaccharide biosynthesis protein Cap5K, putative 3.0 2.1<br />

SAOUHSC_00125 capL cap5L protein/glycosyltransferase, putative 2.8 1.9<br />

SAOUHSC_00126 capM capsular polysaccharide biosynthesis protein Cap8M 2.6 1.8<br />

SAOUHSC_00127 capN cap5N protein/UDP-glucose 4-epimerase, putative 2.7 1.8<br />

SAOUHSC_00128 capO cap5O protein/UDP-N-acetyl-D-mannosaminuronic acid dehydrogenase 2.6 1.7<br />

SAOUHSC_00248 lytM peptidoglycan hydrolase, putative -4.2 -2.9<br />

SAOUHSC_00869 dltA D-alanine-activating enzyme -2.5 -2.0<br />

SAOUHSC_00870 dltB dltB protein, putative -2.6 -2.3<br />

SAOUHSC_00872 dltD extramembranal protein -2.6 -2.3<br />

SAOUHSC_02883 ssaA staphylococcal secretory antigen ssaA -5.5 -3.3<br />

SAOUHSC_03001 icaR ica operon transcriptional regulator IcaR, putative 2.3 1.4<br />

SAOUHSC_03004 icaB icaB protein, putative -3.2 -2.1<br />

SAOUHSC_03005 icaC intercellular adhesion protein C, putative -2.6 -2.2<br />

SAOUHSC_00674 sarX staphylococcal accessory regulator X -4.8 -5.1<br />

SAOUHSC_02799 sarT staphylococcal accessory regulator T -8.8 -8.4<br />

SAOUHSC_02532 sarV staphylococcal accessory regulator V 1.7 2.3<br />

SAOUHSC_02566 sarR staphylococcal accessory regulator R 1.9 2.1<br />

SAOUHSC_02669 sarZ staphylococcal accessory regulator Z 2.0 2.4<br />

a Differentially regulated genes (significant with p* < 0.05 and with a minimal absolute fold change <strong>of</strong> 2) are indicated in bold.<br />

135


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

In vitro Infection Experiment Study<br />

Reproducibility <strong>of</strong> replicates and clustering <strong>of</strong> experimental condition groups<br />

Staphylococcus aureus RN1HG or RN1HG GFP were used to infect S9 cells, a human bronchial<br />

epithelial cell line. The study included samples from the four time points 0 h (start <strong>of</strong><br />

experiment), 1 h, 2.5 h and 6.5 h after start <strong>of</strong> infection, and included besides internalized<br />

samples also control samples from inoculation condition <strong>of</strong> exponential growth phase,<br />

serum/CO 2 controls, non-adherent staphylococci, and an anaerobic incubation control, although<br />

not every sample condition was represented in every time point (see Material and<br />

Methods / In vitro Infection Experiment Study / “Cell culture infection model” and “Bacterial<br />

control samples”, pages 57 and 58). Hierarchical clustering visualized the grouping <strong>of</strong> the array<br />

data sets according to their similarity (Fig. R.5.4). In an upper clustering level, grouping into three<br />

major classes clearly revealed the separation <strong>of</strong> main experimental groups: 1) internalized<br />

staphylococci, 2) controls <strong>of</strong> the later time points 2.5 h and 6.5 h, and 3) exponential growth<br />

phase samples, which represent the inoculation sample condition, and controls <strong>of</strong> the early time<br />

point 1 h (Fig. R.5.4, red dashed line). A lower level <strong>of</strong> clustering separated the individual sample<br />

conditions and time points (Fig. R.5.4, orange dashed line). Here, biological replicates for 2.5 h<br />

internalized, 6.5 h internalized, 2.5 h serum/CO 2 control, 6.5 h serum/CO 2 control, 2.5 h<br />

anaerobic incubation and exponential growth phase samples were arranged together. Samples <strong>of</strong><br />

1 h serum/CO 2 control and 1 h non-adherent staphylococci were an exception: Because they<br />

were very similar, the clustering did not lead to a complete segmentation <strong>of</strong> biological replicates<br />

but additionally grouped 2 samples <strong>of</strong> the two different conditions (Fig. R.5.4, orange dashed line,<br />

lower part). These two control groups were not only similar to each other, but additionally still<br />

featured a high similarity to the inoculation condition <strong>of</strong> exponential growth phase.<br />

S. aureus RN1HG sample conditions and time points<br />

internalized<br />

2.5 h internalized<br />

6.5 h internalized<br />

controls <strong>of</strong><br />

later time<br />

points 2.5 h<br />

and 6.5 h<br />

2.5 h serum/CO 2 control<br />

6.5 h serum/CO 2 control<br />

2.5 h anaerobic incubation<br />

exponential<br />

growth<br />

phase and<br />

controls <strong>of</strong><br />

early time<br />

point 1 h<br />

exponential growth phase<br />

1 h serum/CO 2 control<br />

1 h non-adherent<br />

1 h serum/CO 2 control<br />

1 h non-adherent<br />

Fig. R.5.4: Hierarchical clustering <strong>of</strong> 23 tiling array data sets from in vitro infection experiment.<br />

The following clustering algorithms were applied on z-score-transformed data: Agglomerative clustering with average linkage using<br />

using cosine correlation as similarity measure. All sequences were included in the cluster analysis. The red dashed line indicates<br />

grouping into three major classes: 1) internalized staphylococci, 2) controls <strong>of</strong> the later time points 2.5 h and 6.5 h, and 3) exponential<br />

growth phase samples, which represent the inoculation sample condition, and controls <strong>of</strong> the early time point 1 h. A lower level <strong>of</strong><br />

clustering (orange dashed line) separated the individual sample conditions and time points. Here, biological replicates for 2.5 h<br />

internalized, 6.5 h internalized, 2.5 h serum/CO 2 control, 6.5 h serum/CO 2 control, 2.5 h anaerobic incubation and exponential growth<br />

phase samples were arranged together. Samples <strong>of</strong> 1 h serum/CO 2 control and 1 h non-adherent staphylococci were an exception:<br />

Because they were very similar, the clustering did not lead to a complete segmentation <strong>of</strong> biological replicates, but additionally<br />

grouped 2 samples <strong>of</strong> the two different conditions (orange dashed line, lower part).<br />

136


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

Summing up, the biological replicates exhibited a high reproducibility. Most importantly, the<br />

internalized condition had characteristics very distinct from either the starting condition <strong>of</strong><br />

exponential growth phase and the similar early time point controls, but also from the later<br />

controls which were incubated in the presence <strong>of</strong> serum in 5 % CO 2 -atmosphere without agitation<br />

and from the anaerobic control samples.<br />

Choice <strong>of</strong> adequate baseline samples for comparison <strong>of</strong> the experimental conditions:<br />

Time-matched controls are not suitable because <strong>of</strong> their similarity to stationary growth phase /<br />

stringent response<br />

Many bacterial species are capable <strong>of</strong> inducing a so-called stringent response with the aim to<br />

adapt their metabolism to situations <strong>of</strong> nutrient limitation, especially to carbon and amino acid<br />

starvation. The stringent response leads to induced stress resistance, decelerated growth and<br />

alleviated metabolism. The main mediators <strong>of</strong> the stringent response are pyrophosphorylated<br />

GTP or GDP molecules, which are also called alarmones: guanosine tetraphosphate (ppGpp) and<br />

guanosine pentaphosphate (pppGpp), depending on bacterial species. In the course <strong>of</strong> the<br />

stringent response, up to one third <strong>of</strong> the transcriptome can be modulated. The stringent<br />

response has been intensively studied in E. coli, but also S. aureus is capable <strong>of</strong> such response<br />

(Condon et al. 1995, Crosse et al. 2000, Anderson KL et al. 2006, Wolz et al. 2010).<br />

Entry into stationary growth phase is initiated <strong>by</strong> a beginning nutrient limitation after<br />

consumption in the exponential phase <strong>of</strong> growth. Therefore, in stationary phase bacteria’s<br />

stringent response is triggered.<br />

When comparing the gene expression signature <strong>of</strong> S. aureus RN1HG, which was incubated for<br />

2.5 h in serum-containing infection medium under 5 % CO 2 -atmosphere (37°C), with the signature<br />

<strong>of</strong> stationary phase time point t 2 , which is defined as 2 h after entry into the stationary phase<br />

(each in comparison to exponentially growing bacteria), a high overlap between both signatures<br />

was recognized. Almost 44 % <strong>of</strong> sequences in the 2.5 h serum/CO 2 control signature were also<br />

found in the stationary phase signature (Fig. R.5.5; for S. aureus RN1HG stationary phase<br />

response refer to chapter “Growth Media Comparison Study”, page 131). This first hint for<br />

similarities between the controls <strong>of</strong> later time points in the infection experiment and the<br />

bacterial stationary phase samples provoked a more detailed analysis <strong>of</strong> expression data <strong>of</strong> genes<br />

already known to be involved in the stringent response.<br />

Fig. R.5.5:<br />

Comparison <strong>of</strong> 2.5 h serum/CO 2 control signature<br />

with stationary phase signature.<br />

Each sample condition was compared to its<br />

corresponding baseline samples <strong>of</strong> exponential<br />

growth in log-transformed space using textbook<br />

one-way ANOVA with Benjamini-Hochberg False<br />

Discovery Rate multiple testing correction, and<br />

p* < 0.05 was regarded as significant. An absolute<br />

fold change cut<strong>of</strong>f <strong>of</strong> 2 was applied.<br />

sequences differentially expressed in<br />

the comparison <strong>of</strong> incubation for<br />

2.5 h in serum-containing medium<br />

and 5% CO 2 atmosphere vs.<br />

exponential growth phase in pMEM<br />

487 418 1005<br />

sequences differentially expressed<br />

between stationary growth phase (t 2 )<br />

and exponential growth phase <strong>of</strong><br />

S. aureus RN1HG cultivated in pMEM<br />

Ribosomal proteins are repressed during stringent response (Anderson KL et al. 2006). As the<br />

bacterial metabolism adapts to slower growth and limited energy resources, the cell also<br />

diminishes protein synthesis and therefore does not need to produce further ribosomes.<br />

Reduced expression <strong>of</strong> 55 genes for ribosomal proteins was observed in late serum/CO 2 controls<br />

(2.5 h and 6.5 h) and in anaerobiosis (Fig. R.5.6 A).<br />

137


S. aureus RN1HG sample conditions and time points<br />

mean intensity <strong>of</strong> 2 or 3 biological replicates<br />

S. aureus RN1HG sample conditions and time points<br />

mean intensity <strong>of</strong> 2 or 3 biological replicates<br />

*<br />

*<br />

*<br />

*<br />

S. aureus RN1HG sample conditions and time points<br />

mean intensity <strong>of</strong> 2 or 3 biological replicates<br />

*<br />

*<br />

*<br />

Maren Depke<br />

* p


S. aureus RN1HG sample conditions and time points<br />

mean intensity <strong>of</strong> 2 or 3 biological replicates<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

Similarly, tRNA synthetase genes are repressed in stringent response (Anderson KL et al.<br />

2006). Here, repression <strong>of</strong> these genes was detectable in internalized staphylococci (2.5 h), late<br />

serum/CO 2 controls (2.5 h and 6.5 h) and in samples after 2.5 h anaerobic incubation<br />

(Fig. R.5.6 B). An exception is the isoleucin tRNA synthetase (ileS), which is induced in stringent<br />

response (Anderson KL et al. 2006). But in the 2.5 h and 6.5 h serum/CO 2 controls, this gene<br />

shows a trend <strong>of</strong> repression similar to the other tRNA synthetases with a fold change <strong>of</strong> −1.5 and<br />

−1.8, respectively.<br />

Furthermore, translation elongation factor gene expression is known to be suppressed in<br />

stringent response (Anderson KL et al. 2006). Such repression physiologically acts in concert with<br />

repressed ribosomal protein and tRNA synthetase genes and leads to diminished protein<br />

synthesis. In this study, 4 genes for translation elongation factors revealed a trend <strong>of</strong> repressed<br />

expression in late serum/CO 2 controls (2.5 h and 6.5 h). With p = 0.06, statistical significance was<br />

only slightly missed in the comparison between these two controls and exponential growth<br />

samples (Fig. R.5.6 C).<br />

Stringent response is mediated <strong>by</strong> the regulator RelA whose transcript lateron is increased in<br />

stringent response conditions (Anderson KL et al. 2006). A strong trend <strong>of</strong> induction for relA was<br />

observed for the 2.5 h and 6.5 h serum/CO 2 controls. Although relA exhibited significantly<br />

different expression (p* < 0.01) in the comparison <strong>of</strong> all sequences between the 2.5 h serum/CO 2<br />

control and exponential growth phase samples, the 1.8-fold change did not pass the cut<strong>of</strong>f level<br />

<strong>of</strong> 2. On the other hand, in the comparison between the later 6.5 h serum/CO 2 control and<br />

exponential growth phase samples, the fold change amounted to 2.1, but in this comparison<br />

statistical testing was not possible because <strong>of</strong> limited number <strong>of</strong> replicates relA at the 6.5 h time point<br />

(n = 2). Nevertheless, relA could be regarded as induced in the late serum/CO 2 controls, because<br />

both time points showed the same trend and all six other experimental samples possessed<br />

almost the same, lower gene expression intensity, which proved relA highly reliable measurement <strong>of</strong><br />

the relA expression <strong>by</strong> the tiling array approach (Fig. R.5.7).<br />

Fig. R.5.7:<br />

Comparison <strong>of</strong> mean expression<br />

intensities for the stringent response<br />

regulatory gene relA.<br />

A strong trend <strong>of</strong> relA induction was<br />

visible for the late serum/CO 2 controls<br />

(2.5 h and 6.5 h). Although relA exhibited<br />

significantly different expression (one-way<br />

textbook ANOVA with Benjamini-<br />

Hochberg False Discovery Rate multiple<br />

testing correction and p* < 0.01) in the<br />

comparison between the 2.5 h serum/CO 2<br />

control and exponential growth phase<br />

samples, the 1.8-fold change did not pass<br />

the cut<strong>of</strong>f level <strong>of</strong> 2. On the other hand, in<br />

the comparison between the later 6.5 h<br />

serum/CO 2 control and exponential<br />

growth phase samples, the fold change<br />

amounted to 2.1, but in this comparison<br />

statistical testing was not possible<br />

because <strong>of</strong> limited number <strong>of</strong> replicates at<br />

the 6.5 h time point (n = 2).<br />

exponential growthOD phase 0.4<br />

1 h serum/CO 2 control 1h CO2<br />

1 non-adh. h non-adherent MOI 25<br />

2.5 h internal. internalized 2.5h<br />

6.5 h internal. internalized 6.5h<br />

2.5 h serum/COCO2 2 control 2.5h<br />

6.5 h serum/COCO2 2 control 6.5h<br />

2.5 h anaerobic anaerobic incubation 2.5h<br />

1.0E+04<br />

10000<br />

2.0E+04<br />

20000<br />

mean intensitySamples<br />

<strong>of</strong> biological replicates<br />

3.0E+04<br />

30000<br />

From repressed or in trend repressed ribosomal protein, tRNA synthetase, and translation<br />

elongation factor genes and from the strong tendency <strong>of</strong> relA infection induction experiments it was reasoned medium comparison that the<br />

gene expression signature <strong>of</strong> late serum/CO 2 controls (2.5 h and 6.5 h) in fact resembled the<br />

stationary phase/stringent response. As internalized staphylococci and the early control samples<br />

did not possess this similarity, it was concluded that the 1 h serum/CO 2 control sample, although<br />

not time-matched, was the most appropriate baseline sample for this infection experiment study.<br />

139


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

Comparison <strong>of</strong> experimental condition groups and assessment <strong>of</strong> differentially regulated<br />

genes/sequences<br />

The first general data analysis steps had provided insight into the overall similarity <strong>of</strong> the 23<br />

arrays available after hybridization <strong>of</strong> samples and controls from the infection experiment.<br />

Biological replicates exhibited good reproducibility. The time-matched serum/CO 2 controls for<br />

the internalized samples built a distinct group <strong>of</strong> samples. As a more detailed analysis revealed<br />

similarities <strong>of</strong> these controls with stationary phase/stringent response, it became apparent that<br />

the 1 h serum/CO 2 samples were the most suitable control to use as baseline in this study.<br />

Therefore, the following seven types <strong>of</strong> comparisons were chosen to elucidate the changes in<br />

gene expression pattern and lateron the physiological reactions <strong>of</strong> S. aureus RN1HG in the<br />

settings <strong>of</strong> the in vitro S9 infection model (Fig. R.5.8):<br />

<br />

<br />

<br />

<br />

<br />

<br />

<br />

comparison <strong>of</strong> exponential growth phase samples, which represent the inoculation<br />

condition, with baseline <strong>of</strong> 1 h serum/CO 2 controls<br />

comparison <strong>of</strong> non-adherent staphylococci (1 h) with baseline <strong>of</strong> 1 h serum/CO 2 controls<br />

comparison <strong>of</strong> 2.5 h internalized samples with baseline <strong>of</strong> 1 h serum/CO 2 controls<br />

comparison <strong>of</strong> 6.5 h internalized samples with baseline <strong>of</strong> 1 h serum/CO 2 controls<br />

comparison <strong>of</strong> 2.5 h serum/CO 2 controls with baseline <strong>of</strong> 1 h serum/CO 2 controls<br />

comparison <strong>of</strong> 6.5 h serum/CO 2 controls with baseline <strong>of</strong> 1 h serum/CO 2 controls<br />

comparison <strong>of</strong> 2.5 h anaerobic incubation samples with baseline <strong>of</strong> 1 h serum/CO 2 controls<br />

inoculation<br />

condition<br />

and early<br />

time point<br />

control<br />

samples<br />

1<br />

1 h<br />

serum/CO 2<br />

control<br />

(3 arrays)<br />

exponential<br />

growth phase<br />

(3 arrays)<br />

2<br />

1h<br />

non-adherent<br />

S. aureus<br />

(3 arrays)<br />

4<br />

3<br />

7<br />

5<br />

6<br />

6.5 h<br />

serum/CO 2<br />

control<br />

(2 arrays)<br />

2.5 h<br />

internalized<br />

S. aureus<br />

(3 arrays)<br />

2.5 h<br />

serum/CO 2<br />

control<br />

(3 arrays)<br />

6.5 h<br />

internalized<br />

S. aureus<br />

(3 arrays)<br />

internalized<br />

samples<br />

2.5 h<br />

anaerobic<br />

incubation<br />

(3 arrays)<br />

late<br />

time point<br />

controls<br />

Fig. R.5.8: Overview on the comparisons between groups in this study that were partly addressed with statistical testing and visualized<br />

with scatter plots.<br />

Baseline for all comparisons was the serum/CO 2 control group <strong>of</strong> 1 h. The very similar exponential growth phase samples, which<br />

represent the inoculation condition, were compared to this baseline () and the even more similar non-adherent staphylococcal<br />

samples after 1 h (). To elucidate the gene expression signature <strong>of</strong> internalized bacteria, the 2.5 h () as well as the 6.5 h ()<br />

internalized samples were compared to the samples <strong>of</strong> 1 h serum/CO 2 control group. For comparison purposes, the late time point<br />

controls were also checked against the early serum/CO 2 control as baseline: 2.5 h serum/CO 2 controls () and samples after 2.5 h <strong>of</strong><br />

anaerobic incubation (). The comparison focusing on the difference between 6.5 h and 1 h serum/CO 2 control samples () could<br />

not be performed with statistical testing because <strong>of</strong> the low number (n = 2) <strong>of</strong> arrays at the 6.5 h time point.<br />

140


exponential growth phase<br />

2.5 h internalized<br />

2.5 h anaerobic incubation<br />

2.5 h serum/CO 2 control<br />

1 h non-adherent<br />

6.5 h internalized<br />

6.5 h serum/CO 2 control<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

The comparisons were visualized using scatter plots (Fig. R.5.9). Comparable to the clustering<br />

results, a high concordance between expression values <strong>of</strong> 1 h serum/CO 2 controls and<br />

exponential phase samples, which represent the inoculation condition (Fig. R.5.9, panel ), and<br />

even more striking between 1 h serum/CO 2 controls and non-adherent staphylococcal samples<br />

after 1 h (Fig. R.5.9, panel ) was observed. Strong effects <strong>of</strong> treatment in comparison to 1 h<br />

serum/CO 2 controls emerged in 2.5 h internalized (Fig. R.5.9, panel ), 6.5 h internalized<br />

(Fig. R.5.9, panel ), 2.5 h serum/CO 2 controls (Fig. R.5.9, panel ), 6.5 h serum/CO 2 controls<br />

(Fig. R.5.9, panel ), and in 2.5 h anaerobically incubated samples (Fig. R.5.9, panel ). Most<br />

variation was detectable in the comparison <strong>of</strong> 1 h and 6.5 h serum/CO 2 controls (Fig. R.5.9,<br />

panel ), but this can be due to the lower number <strong>of</strong> only two replicates for the 6.5 h serum/CO 2<br />

control, which led to a lesser compensation for variation in the mean value than in the mean <strong>of</strong><br />

three replicates in the other sample condition/time point groups.<br />

<br />

<br />

1 h serum/CO 2 control<br />

1 h serum/CO 2 control<br />

<br />

<br />

1 h serum/CO 2 control<br />

1 h serum/CO 2 control<br />

<br />

<br />

1 h serum/CO 2 control<br />

1 h serum/CO 2 control<br />

<br />

1 h serum/CO 2 control<br />

Fig. R.5.9: Scatter plots comparing mean signal intensities<br />

<strong>of</strong> treatment groups.<br />

The signals <strong>of</strong> the eight groups <strong>of</strong> “1 h serum/CO 2<br />

controls”, “exponential growth phase”, “1 h nonadherent<br />

staphylococci”, “2.5 h internalization”, “6.5 h<br />

internalization”, “2.5 h serum/CO 2 controls”, “6.5 h<br />

serum/CO 2 controls”, and “2.5 h anaerobic incubation”<br />

were plotted after combining the three biological<br />

replicates. In case <strong>of</strong> “6.5 h serum/CO 2 controls” only two<br />

biological replicates were available. All sequences<br />

available on the array are shown. Numbers in the first<br />

column refer to the comparison scheme in Fig. R.5.8.<br />

141


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

The different experimental groups were compared with statistical testing to obtain lists <strong>of</strong><br />

differentially expressed sequences (Table R.5.4). It has been already described that the three<br />

groups exponential growth phase as inoculation condition, 1 h serum/CO 2 control, and 1 h nonadherent<br />

staphylococci were similar to each other. Non-surprisingly, when comparing 1 h nonadherent<br />

staphylococci to the 1 h serum/CO 2 control samples, no statistically significant<br />

differential expression was detected. Between exponential growth phase and 1 h serum/CO 2<br />

control, 156 sequences were differentially expressed <strong>of</strong> which 76 exceeded a minimal absolute<br />

fold change <strong>of</strong> 2. In the comparison <strong>of</strong> 2.5 h and 6.5 h internalization vs. 1 h serum/CO 2 control,<br />

1392 and 1205 sequences were regarded as differentially expressed, <strong>of</strong> which 765 and 627<br />

possessed a minimal absolute fold change <strong>of</strong> 2, respectively. Additionally, 1749 sequences were<br />

differentially expressed between 2.5 h serum/CO 2 control and 1 h serum/CO 2 control, and 905 <strong>of</strong><br />

these passed the minimal absolute fold change cut<strong>of</strong>f 2. The 2.5 h anaerobic signature in<br />

comparison with the baseline <strong>of</strong> 1 h serum/CO 2 control included 1339 sequences, and still 680<br />

exhibited a minimal absolute fold change <strong>of</strong> 2.<br />

Table R.5.4: Results <strong>of</strong> group comparisons with statistical testing <strong>of</strong> S9 infection experiment staphylococcal array data sets.<br />

group comparison a<br />

number <strong>of</strong> sequences<br />

significant with p* < 0.05 in<br />

textbook one-way ANOVA with<br />

Benjamini-Hochberg False Discovery<br />

Rate multiple testing correction<br />

absolute fold change equal<br />

to or greater than 2 AND<br />

significant with p* < 0.05<br />

exponential growth phase vs. 1 h serum/CO 2 control 156 76<br />

1 h non-adherent staphylococci vs. 1 h serum/CO 2 control 0 -<br />

2.5 h internalization vs. 1 h serum/CO 2 control 1392 765<br />

6.5 h internalization vs. 1 h serum/CO 2 control 1205 627<br />

2.5 h serum/CO 2 control vs. 1 h serum/CO 2 control 1749 905<br />

2.5 h anaerobic incubation vs. 1 h serum/CO 2 control 1339 680<br />

a The 080604_SA_JH_Tiling array contains 3882 sequences <strong>of</strong> which 2825 are known transcripts. The remaining 1057 sequences were<br />

identified as new transcripts in the tiling array analysis. All sequences were included in statistical testing.<br />

Fractions <strong>of</strong> known genes and newly identified transcribed sequences included on the tiling<br />

array and in the lists <strong>of</strong> differentially expressed sequences derived from the comparisons <strong>of</strong><br />

infection experiment samples<br />

The 080604_SA_JH_Tiling array contains 3882 sequences <strong>of</strong> which 2825, corresponding to<br />

73 %, are known transcripts and associated with a LocusTag number (SAOUHSC_*). The remaining<br />

1057 sequences were identified as new transcripts in the tiling array analysis. These new<br />

transcripts are expected to belong to all types <strong>of</strong> RNAs like mRNA and regulatory RNA, but also<br />

formerly unknown transcribed fragments 5’ and 3’ to known genes might be included.<br />

Furthermore, some artifacts might still be included. All sequences, known and new, were<br />

included in statistical testing.<br />

For the internalization-specific signature, 200 and 138 differentially expressed transcripts<br />

belonged to the group <strong>of</strong> new transcripts at the 2.5 h and 6.5 h time point, respectively<br />

(Table R.5.5). Furthermore, 19 new transcripts were included in the set <strong>of</strong> differentially expressed<br />

sequences between the inoculation condition <strong>of</strong> exponential growth and the baseline <strong>of</strong> 1 h<br />

serum/CO 2 control. In comparison to that baseline, 2.5 h serum/CO 2 control and 2.5 h<br />

anaerobically incubated samples inclosed 250 and 200 differentially expressed new transcripts,<br />

respectively (Table R.5.5).<br />

142


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

Table R.5.5: Fractions <strong>of</strong> known annotated and newly detected transcripts in the results <strong>of</strong> group comparisons with statistical testing<br />

<strong>of</strong> S9 infection experiment staphylococcal array data sets.<br />

group comparison a<br />

total number <strong>of</strong><br />

sequences with<br />

absolute fold<br />

change equal to or<br />

greater than 2 AND<br />

significant with<br />

p* < 0.05<br />

number <strong>of</strong> known<br />

transcribed<br />

sequences with<br />

absolute fold<br />

change equal to or<br />

greater than 2 AND<br />

significant with<br />

p* < 0.05<br />

number <strong>of</strong> new<br />

transcribed sequences<br />

with absolute fold<br />

change equal to or<br />

greater than 2 AND<br />

significant with<br />

p* < 0.05<br />

exponential growth phase vs. 1 h serum/CO 2 control 76 57 19<br />

2.5 h internalization vs. 1 h serum/CO 2 control 765 565 200<br />

6.5 h internalization vs. 1 h serum/CO 2 control 627 489 138<br />

2.5 h serum/CO 2 control vs. 1 h serum/CO 2 control 905 655 250<br />

2.5 h anaerobic incubation vs. 1 h serum/CO 2 control 680 480 200<br />

a The 080604_SA_JH_Tiling array contains 3882 sequences <strong>of</strong> which 2825 are known transcripts. The remaining 1057 sequences were<br />

identified as new transcripts in the tiling array analysis. All sequences were included in statistical testing.<br />

Comparison <strong>of</strong> differentially regulated sequences in 2.5 h and 6.5 h internalized staphylococci<br />

The gene expression signatures <strong>of</strong> 2.5 h and 6.5 h internalized staphylococci, which resulted<br />

from statistical comparison with the baseline <strong>of</strong> 1 h serum/CO 2 control including multiple testing<br />

correction and minimal absolute fold change cut<strong>of</strong>f 2, were compared in a Venn diagram<br />

(Fig. R.5.10 A). For 408 sequences, differential expression was observed in both time points <strong>of</strong><br />

internalized samples, while 357 and 219 sequences were specific for the 2.5 h and 6.5 h<br />

internalization signature, respectively. This corresponds to a fraction <strong>of</strong> 53 % <strong>of</strong> sequences<br />

regulated after 2.5 h <strong>of</strong> internalization, which were also differentially expressed after 6.5 h. Vice<br />

versa, 65 % <strong>of</strong> differentially expressed sequences at the 6.5 h time point were also found to be<br />

regulated after 2.5 h. When examining the induced and repressed sequences separately, it was<br />

first obvious that sequences regulated at both time points always possessed the same regulation<br />

direction in these two time points (Fig. R.5.10 B). Further, induction accounted for a bigger<br />

fraction than reduction in the differentially expressed sequences, although the excess was only<br />

small at the 6.5 h time point.<br />

A<br />

B<br />

357 408 219<br />

2.5 h<br />

repressed<br />

repressed 6.5 h<br />

induced<br />

176 121<br />

induced<br />

181<br />

187<br />

98<br />

sequences differentially<br />

expressed in the comparison<br />

“2.5 h internalization” vs.<br />

“1 h serum/CO 2 control”<br />

sequences differentially<br />

expressed in the comparison<br />

“6.5 h internalization” vs.<br />

“1 h serum/CO 2 control”<br />

221<br />

Fig. R.5.10: Comparison <strong>of</strong> the 2.5 h and 6.5 h signatures <strong>of</strong> internalized staphylococci.<br />

Both internalized samples were compared to the baseline <strong>of</strong> 1 h serum/CO 2 control with statistical testing and multiple testing<br />

correction (p* < 0.05), and a minimal absolute fold change cut<strong>of</strong>f <strong>of</strong> 2 was applied. The comparison <strong>of</strong> differentially expressed<br />

sequences at the 2.5 h and 6.5 h time point was performed for all regulated sequences (A) and for induced and repressed sequences<br />

separately (B).<br />

In the 2.5 h samples 402 sequences were induced (181 specifically for that time point), 363<br />

were repressed (176 specifically for that time point), while in the 6.5 h samples 319 exhibited<br />

induction (98 specifically for that time point) and 308 exhibited repression (121 specifically for<br />

that time point). Therefore, the 408 sequences differentially expressed in both time points<br />

consisted <strong>of</strong> 221 induced and 187 repressed sequences.<br />

143


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

Comparison <strong>of</strong> differentially expressed genes [Maren Depke] and proteins with differing<br />

abundance [Sandra Scharf] in internalized staphylococci<br />

In an experimental setup corresponding to the generation <strong>of</strong> samples for tiling array analysis,<br />

the abundance <strong>of</strong> proteins in internalized staphylococci was recorded with a combined approach<br />

<strong>of</strong> stable isotope labeling with amino acids in cell culture (SILAC), FACS cell sorting and mass<br />

spectrometric analysis (Sandra Scharf, Schmidt et al. 2010). The method, which was applied for<br />

this first proteome pr<strong>of</strong>iling <strong>of</strong> in vitro internalized staphylococci, limits the possible protein<br />

identification mainly to cytosolic proteins: Secreted proteins were lost during sample<br />

preparation, and peptides from membrane associated proteins are difficult to obtain <strong>by</strong> tryptic<br />

digest <strong>of</strong> whole-cell samples.<br />

In the proteomic study, the time points 1.5 h, 2.5 h, 3.5 h, 4.5 h, 5.5 h, and 6.5 h after start <strong>of</strong><br />

infection were included, i. e. the analysis window was divided into smaller sections than in the<br />

tiling array study. In an analysis which included more samples and controls than published in the<br />

pilot study, Sandra Scharf identified 648 proteins and obtained a set <strong>of</strong> 114 proteins with<br />

differential abundance. Differential abundance was assigned to proteins when the regulation was<br />

observed in at least 2 <strong>of</strong> 3 biological replicates in at least 1 <strong>of</strong> 6 analyzed time points, and<br />

proteins with contradictory results were excluded. The regulated proteins were normalized in<br />

reference to the SILAC-method immanent internal control (heavy isotope labeled peptides) and<br />

exhibited an abundance difference when comparing target peptide ratios with the median ratio<br />

for all peptides at the corresponding time point. Regulated proteins which were additionally<br />

different in abundance in a serum/CO 2 control (without presence <strong>of</strong> <strong>host</strong> cells) were excluded.<br />

To gain an insight into the comparability <strong>of</strong> transcriptome and proteome pr<strong>of</strong>iling,<br />

differentially expressed sequences were compared to regulated proteins. First, up-regulated<br />

proteins were compared to the sequences which exhibited increased expression 2.5 h or 6.5 h<br />

after start <strong>of</strong> infection or in both time points (Fig. R.5.11 A). Second, the same comparison was<br />

performed for down-regulated proteins and repressed sequences (Fig. R.5.11 B).<br />

A<br />

B<br />

up-regulated proteins in<br />

internalized staphylococci<br />

down-regulated proteins in<br />

internalized staphylococci<br />

53<br />

34<br />

5 2<br />

14<br />

176 96<br />

207<br />

3 0<br />

6<br />

173 121<br />

181<br />

induced sequences in<br />

“2.5 h internalization” vs.<br />

“1 h serum/CO 2 control”<br />

induced sequences in<br />

“6.5 h internalization” vs.<br />

“1 h serum/CO 2 control”<br />

repressed sequences in<br />

“2.5 h internalization” vs.<br />

“1 h serum/CO 2 control”<br />

repressed sequences in<br />

“6.5 h internalization” vs.<br />

“1 h serum/CO 2 control”<br />

Fig. R.5.11: Comparison <strong>of</strong> differentially expressed sequences in at least one <strong>of</strong> the two analyzed time points with proteins exhibiting<br />

different abundance in internalized staphylococci.<br />

A. Up-regulated proteins and induced sequences. B. Down-regulated proteins and repressed sequences.<br />

144


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

In total, the transcripts <strong>of</strong> 21 up-regulated proteins were increased in at least one <strong>of</strong> the two<br />

analyzed time points. These consisted <strong>of</strong> 14 proteins, whose genes were induced at both time<br />

points, and <strong>of</strong> 5 and 2 proteins, whose gene expression was increased only at the 2.5 h and 6.5 h<br />

time point, respectively. For 6 proteins, whose abundance was reduced, the corresponding gene<br />

expression was repressed at both time points, and for further 3 down-regulated proteins the<br />

gene expression was only repressed at the 2.5 h time point (Table R.5.6). These numbers result in<br />

an overlap <strong>of</strong> 28 % <strong>of</strong> up- and 21 % <strong>of</strong> down-regulated proteins with the corresponding gene<br />

expression changes. As the lists <strong>of</strong> differentially expressed sequences included newly identified<br />

transcripts from the tiling array, which were not accessed with mass spectrometry identification,<br />

these lists contained many transcriptome-specific results.<br />

Table R.5.6: Differentially expressed sequences in internalization samples which exhibit differences in protein abundance and<br />

regulation in the same direction.<br />

LocusTag gene annotation<br />

difference<br />

in protein<br />

abundance<br />

a<br />

fold change in comparison<br />

to baseline<br />

1 h serum/CO 2 control<br />

2.5 h<br />

internalized b<br />

6.5 h<br />

internalized b<br />

SAOUHSC_01846 acsA acetyl-CoA synthetase, putative + 16.3 4.0<br />

SAOUHSC_01395 asd aspartate-semialdehyde dehydrogenase + 14.1 13.3<br />

SAOUHSC_00086 butA 3-ketoacyl-acyl carrier protein reductase, putative + 2.5 2.2<br />

SAOUHSC_01396 dapA dihydrodipicolinate synthase + 12.5 12.0<br />

SAOUHSC_01398 dapD<br />

2,3,4,5-tetrahydropyridine-2-carboxylate N-<br />

succinyltransferase, putative<br />

+ 9.1 9.8<br />

SAOUHSC_01320 dhoM homoserine dehydrogenase, putative + 2.1 5.4<br />

SAOUHSC_00196 fadB hypothetical protein SAOUHSC_00196 + 84.4 16.6<br />

SAOUHSC_00197 fadD hypothetical protein SAOUHSC_00197 + 100.0 35.8<br />

SAOUHSC_02869 rocA delta-1-pyrroline-5-carboxylate dehydrogenase, putative + 8.0 2.8<br />

SAOUHSC_01418 odhA 2-oxoglutarate dehydrogenase, E1 component + 5.2 2.2<br />

SAOUHSC_01416 odhB<br />

2-oxoglutarate dehydrogenase, E2 component,<br />

dihydrolipoamide succinyltransferase<br />

+ 4.5 2.1<br />

SAOUHSC_00371 hypothetical protein SAOUHSC_00371 + 4.4 2.1<br />

SAOUHSC_01399 hypothetical protein SAOUHSC_01399 + 7.9 8.3<br />

SAOUHSC_02425 hypothetical protein SAOUHSC_02425 + 5.3 2.8<br />

SAOUHSC_01276 glpK glycerol kinase + 2.7 1.7<br />

SAOUHSC_01801 icd, citC isocitrate dehydrogenase, NADP-dependent + 2.1 1.2<br />

SAOUHSC_01972 prsA protein export protein PrsA, putative + 2.0 1.6<br />

SAOUHSC_00767 yfiA hypothetical protein SAOUHSC_00767 + 2.9 1.7<br />

SAOUHSC_00951 hypothetical protein SAOUHSC_00951 + 2.4 1.6<br />

SAOUHSC_01321 thrC threonine synthase + 1.7 5.3<br />

SAOUHSC_00833 hypothetical protein SAOUHSC_00833 + 1.2 2.1<br />

SAOUHSC_01771 hemL glutamate-1-semialdehyde-2,1-aminomutase − -2.2 -2.5<br />

SAOUHSC_01092 pheS phenylalanyl-tRNA synthetase, alpha subunit − -2.5 -3.4<br />

SAOUHSC_01093 pheT phenylalanyl-tRNA synthetase, beta subunit − -2.7 -3.5<br />

SAOUHSC_01010 purC<br />

phosphoribosylaminoimidazole-succinocarboxamide<br />

synthase<br />

− -4.5 -2.2<br />

SAOUHSC_01011 purS<br />

phosphoribosylformylglycinamidine synthase, PurS<br />

protein<br />

− -4.9 -2.3<br />

SAOUHSC_01788 thrS threonyl-tRNA synthetase − -2.1 -2.4<br />

SAOUHSC_01018 purD phosphoribosylamine-glycine ligase − -2.6 -1.1<br />

SAOUHSC_01013 purL phosphoribosylformylglycinamidine synthase II − -3.4 -1.5<br />

SAOUHSC_01012 purQ phosphoribosylformylglycinamidine synthase I − -4.0 -1.9<br />

a Up-regulated protein abundance in internalized staphylococci is indicated <strong>by</strong> “+”, down-regulated abundance <strong>by</strong> ”–“.<br />

b Differentially regulated genes (significant with p* < 0.05 and with a minimal absolute fold change <strong>of</strong> 2) are indicated in bold.<br />

145


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

In a similar comparison, up-regulated proteins were compared with repressed transcripts and<br />

vice versa down-regulated proteins with induced transcripts, which resulted in an overlap <strong>of</strong> 5<br />

(rocD, ldh2, sodM, fabH, SAOUHSC_02150) and 9 (sufC, sufD, spoVG, dps, SAOUHSC_00533,<br />

SAOUHSC_02443, SAOUHSC_02363, SAOUHSC_01854, SAOUHSC_00845) proteins, respectively.<br />

Such contradictory results could be explained in most cases <strong>by</strong> the following characteristics:<br />

transcript changes do not necessarily influence the protein abundance at the same time<br />

point; the potential time shift between both effects is not defined and may vary between<br />

the different proteins<br />

effects on protein or expression might be outside the analysis window (between start <strong>of</strong><br />

infection and first analysis time point <strong>of</strong> 1.5 h or 2.5 h; after last analysis time point <strong>of</strong><br />

6.5 h; between analysis time points especially in the transcriptome analysis)<br />

special protein effects like temporary new synthesis, increased turnover, and other nontranscriptionally<br />

mediated effects<br />

special protein analysis effects: early changes or changes occurring before the first analysis<br />

time point, which stop in the middle <strong>of</strong> the observation period, might still lead to an<br />

absolute fold change value exceeding 2 at the later time points in cases <strong>of</strong> stable proteins<br />

protein analysis challenges like low intensity measurements, high variation between<br />

biological replicates, missing data for one <strong>of</strong> the biological replicates<br />

Expression <strong>of</strong> genes containing binding sites for Rex, an anaerobiosis regulator and redox<br />

sensor<br />

During in vitro infection/internalization experiments, staphylococci are subjected to changes<br />

in oxygen availability. First, they are shifted from aerobic shaking cultures into cell culture dishes,<br />

second, they are incubated without agitation in a CO 2 -atmosphere optimal for the <strong>host</strong> cells, and<br />

finally, internalization into eukaryotic <strong>host</strong> cells might further influence the oxygen availability. To<br />

answer the question whether oxygen limitation has major impact on internalized staphylococci,<br />

the gene expression <strong>of</strong> anaerobiosis-related genes was analyzed. Very recently, Pagels et al. have<br />

published transcriptome and proteome data <strong>of</strong> an anaerobic gene expression regulon (Pagels et<br />

al. 2010). Although indirect effects <strong>of</strong> Rex have also been described, the central regulator Rex, a<br />

redox sensor, acts as transcriptional repressor. In situations <strong>of</strong> increasing NADH, DNA-binding <strong>of</strong><br />

Rex is inhibited, resulting in de-repression <strong>of</strong> genes which possess a Rex binding site. Pagels et al.<br />

defined 21 genes with such binding site, 19 <strong>of</strong> which could be mapped to S. aureus NCTC8325<br />

LocusTags (NCBI’s Entrez Genome Gene Plot; http://www.ncbi.nlm.nih.gov/sutils/geneplot.cgi).<br />

For a first impression, the fraction <strong>of</strong> these genes which were differentially expressed in<br />

staphylococci after 2.5 h <strong>of</strong> anaerobic incubation, but also in 2.5 h and 6.5 h internalized<br />

staphylococci was determined. In anaerobically incubated staphylococci, 5 Rex binding site<br />

containing genes were found to be regulated (Fig. R.5.12 A), whereas 10 <strong>of</strong> these genes were<br />

regulated in internalization at both time points and further 3 genes specifically at the time point<br />

<strong>of</strong> 6.5 h after start <strong>of</strong> infection (Fig. R.5.12 B). A more detailed analysis revealed that while<br />

induction <strong>of</strong> these genes in anaerobically incubated staphylococci was expected, the direction <strong>of</strong><br />

regulation in internalized staphylococci was clearly opposite (Fig. R.5.12 C). Therefore,<br />

internalization probably did not result in stimuli which were able to inactivate the Rex repressor.<br />

146


mean normalized intensity<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

A<br />

14 5 675<br />

B<br />

genes containing Rex binding<br />

sites in S. aureus according to<br />

Pagels et al. 2010<br />

genes containing Rex binding<br />

sites in S. aureus according to<br />

Pagels et al. 2010<br />

sequences differentially<br />

expressed in the comparison<br />

“2.5 h anaerobic incubation”<br />

vs. “1 h serum/CO 2 control”<br />

Fig. R.5.12:<br />

Expression <strong>of</strong> genes containing Rex binding sites.<br />

A. Comparison <strong>of</strong> genes with Rex binding sites (according to<br />

Pagels et al. 2010) and S. aureus RN1HG anaerobic gene<br />

expression signature in the infection experiment study.<br />

B. Comparison <strong>of</strong> genes with Rex binding sites (according to<br />

Pagels et al. 2010) and S. aureus RN1HG S9 internalization gene<br />

expression signature 2.5 h and 6.5 h after start <strong>of</strong> infection.<br />

C. Overview on mean normalized gene expression intensity for<br />

19 genes containing Rex binding sites. After the global<br />

normalization <strong>by</strong> inter-chip scaling and detrending, each<br />

individual gene has been normalized to the expression level <strong>of</strong><br />

the baseline sample “1 h serum/CO 2 control”. Although not<br />

being continuous data, values were depicted as line graphs<br />

instead <strong>of</strong> bar charts for better facility <strong>of</strong> inspection. Black<br />

indicates 8 genes repressed in internalized staphylococci at the<br />

2.5 h or 6.5 h time point, dark gray 5 genes differentially<br />

expressed in at least one internalization time point and after<br />

anaerobic incubation, and light gray marks the remaining 6<br />

genes <strong>of</strong> which 3 possess a fold change value > 1.5 in the<br />

anaerobic sample compared to baseline.<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2<br />

control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h<br />

internalization; 2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h co. −<br />

6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic<br />

incubation)<br />

C<br />

differentially expressed<br />

sequences in “2.5 h<br />

internalization” vs.<br />

“1 h serum/CO 2 control”<br />

10E+01 10.000<br />

10E+00 1.000<br />

10E-01 0.100<br />

10E-02 0.010<br />

10E-03 0.001<br />

exp.<br />

6<br />

0 3<br />

10<br />

357 216<br />

398<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

differentially expressed<br />

sequences in “6.5 h<br />

internalization” vs.<br />

“1 h serum/CO 2 control”<br />

OD 0.4 CO2 (serum) internalized internalized CO2 (serum) CO2 (serum) anaerobic<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

0 h 1 h 2.5 h 6.5 h 2.5 h 6.5 h 2.5 h<br />

S. aureus RN1HG sample conditions and time points<br />

Genes <strong>of</strong> the SaeRS regulon exhibiting differential expression in internalized staphylococci<br />

The two-component system SaeRS is known to positively regulate different virulence<br />

associated genes, e. g. genes coding for adhesins or toxins. The induction <strong>of</strong> gene expression <strong>of</strong><br />

saeR was observed in internalized staphylococci 6.5 h after start <strong>of</strong> infection. Additionally, the<br />

gene saeS for the histidine kinase sensor component was significantly different in 6.5 h<br />

internalized staphylococci compared to baseline, but the fold change <strong>of</strong> 1.7 did not meet the<br />

cut<strong>of</strong>f criterium <strong>of</strong> 2 (Fig. R.5.13 A). As the the two-component system SaeRS is known to be<br />

subjected to positive autoinduction, the induction <strong>of</strong> saeR and trend for saeS provoked a detailed<br />

analysis <strong>of</strong> genes known to be regulated <strong>by</strong> this system. In 2006, Rogasch et al. have published<br />

transcriptome and secretome data <strong>of</strong> the SaeRS regulon in S. aureus strains COL and Newman<br />

(Rogasch et al. 2006). Using a global microarray screening <strong>of</strong> saeRS mutants and parental strains,<br />

Rogasch et al. defined 44 genes which were induced <strong>by</strong> the SaeRS two-component system, 40 <strong>of</strong><br />

which could be mapped to S. aureus NCTC8325 LocusTags (NCBI’s EntrezGenome Gene Plot;<br />

http://www.ncbi.nlm.nih.gov/sutils/geneplot.cgi). For a first impression, the fraction <strong>of</strong> these<br />

genes which were differentially expressed in 2.5 h and 6.5 h internalized staphylococci was<br />

determined (Fig. R.5.13 B). At the 2.5 h time point, 4 genes with known regulation <strong>by</strong> SaeRS were<br />

147


mean normalized intensity<br />

mean normalized intensity<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

differentially expressed, 10 further genes were found to be regulated specifically after 6.5 h,<br />

whereas another 10 genes were regulated in both internalization time points. Of these 4, 10, and<br />

10 genes, the numbers <strong>of</strong> 3, 9, and 9 genes were increased in expression. The remaining gene <strong>of</strong><br />

each group exhibited repression (Fig. R.5.13 C, Table R.5.7).<br />

A<br />

33<br />

B<br />

genes regulated <strong>by</strong> SaeRS<br />

according to Rogasch et al. 2006<br />

22<br />

16<br />

11<br />

saeR<br />

saeS<br />

00<br />

OD 0.4 CO2 (serum) internalized internalized CO2 (serum) CO2 (serum) anaerobic<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

0 h 1 h 2.5 h 6.5 h 2.5 h 6.5 h 2.5 h<br />

S. aureus RN1HG sample conditions and time points<br />

differentially expressed<br />

sequences in “2.5 h<br />

internalization” vs.<br />

“1 h serum/CO 2 control”<br />

4 10<br />

10<br />

353 209<br />

398<br />

differentially expressed<br />

sequences in “6.5 h<br />

internalization” vs.<br />

“1 h serum/CO 2 control”<br />

Fig. R.5.13:<br />

Expression <strong>of</strong> genes known to be regulated <strong>by</strong> SaeRS.<br />

A. Expression <strong>of</strong> saeR and saeS in S. aureus RN1HG during<br />

the S9 infection experiment. Induction <strong>of</strong> both genes was<br />

significant in statistical testing for the 6.5 h time point <strong>of</strong><br />

internalization, but did not pass the two-fold cut<strong>of</strong>f for<br />

saeS.<br />

B. Comparison <strong>of</strong> genes known to be regulated <strong>by</strong> SaeRS<br />

(according to Rogasch et al. 2006) and S. aureus RN1HG S9<br />

internalization gene expression signature 2.5 h and 6.5 h<br />

after start <strong>of</strong> infection.<br />

C. Overview on mean normalized gene expression intensity<br />

for 24 genes known to be regulated <strong>by</strong> SaeRS. After the<br />

global normalization <strong>by</strong> inter-chip scaling and detrending,<br />

each individual gene has been normalized to the<br />

expression level <strong>of</strong> the baseline sample “1 h serum/CO 2<br />

control”. Although not being continuous data, values were<br />

depicted as line graphs instead <strong>of</strong> bar charts for better<br />

facility <strong>of</strong> inspection. Black indicates 9 genes induced in<br />

internalized staphylococci at the 2.5 h or 6.5 h time point,<br />

dark gray 9 genes induced at 6.5 h and 3 genes induced at<br />

2.5 h after start <strong>of</strong> infection, and light gray marks the<br />

remaining 3 genes <strong>of</strong> which one was repressed at both<br />

time points, another specifically 2.5 h and the third<br />

specifically 6.5 h after start <strong>of</strong> infection.<br />

C<br />

10E+02 100.000<br />

10E+01 10.000<br />

10E+00 1.000<br />

10E-01 0.100<br />

10E-02 0.010<br />

OD 0.4 CO2 (serum) internalized internalized CO2 (serum) CO2 (serum) anaerobic<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

0 h 1 h 2.5 h 6.5 h 2.5 h 6.5 h 2.5 h<br />

S. aureus RN1HG sample conditions and time points<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2 control;<br />

2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h internalization;<br />

2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h co. − 6.5 h serum/CO 2<br />

control; 2.5 h anae. − 2.5 h anaerobic incubation)<br />

The set <strong>of</strong> 21 SaeRS regulated genes, which exhibited differential expression in at least one <strong>of</strong><br />

the two analyzed internalization time points, included certain functional groups <strong>of</strong> virulence<br />

factors, and some could even be assigned to more than one group. Membrane bound<br />

adhesins/MSCRAMMS were represented <strong>by</strong> the fibrinogen binding proteins A and B (fnbA, fnbB),<br />

soluble adhesins/SERAMs were covered e. g. <strong>by</strong> extracellular adherence protein (eap), toxins<br />

were included with the examples <strong>of</strong> different hemolysins, chemotaxis inhibitory protein (chp)<br />

served as example for immune-evasive proteins, and finally, also secreted enzymes like serine<br />

protease (spl) were listed. Thus, further data mining was performed in consideration <strong>of</strong> functional<br />

groups <strong>of</strong> virulence factors.<br />

148


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

Table R.5.7: Expression <strong>of</strong> genes known to be regulated <strong>by</strong> SaeRS (according to Rogasch et al. 2006), which exhibited differential<br />

expression in internalized S. aureus RN1HG in S9 cells.<br />

LocusTag gene annotation<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

2.5 h<br />

internalized a<br />

6.5 h<br />

internalized a<br />

SAOUHSC_01942 splA serine protease SplA 3.2 30.4<br />

SAOUHSC_01939 splC serine protease SplC 4.8 43.1<br />

SAOUHSC_02803 fnbA fibronectin-binding protein precursor, putative 6.2 6.4<br />

SAOUHSC_02802 fnbB fibronectin binding protein B, putative 7.0 3.1<br />

SAOUHSC_02709 hlgC leukocidin s subunit precursor, putative 5.3 17.9<br />

SAOUHSC_02708 gamma-hemolysin h-gamma-ii subunit, putative 9.6 19.4<br />

SAOUHSC_00196 fadB hypothetical protein SAOUHSC_00196 84.4 16.6<br />

SAOUHSC_00232 lrgA antiholin-like protein lrgA 4.0 5.3<br />

SAOUHSC_01921 hypothetical protein SAOUHSC_01921 4.8 2.2<br />

SAOUHSC_02710 hlgB leukocidin f subunit precursor 3.4 13.4<br />

SAOUHSC_01121 hla alpha-hemolysin precursor 1.9 3.2<br />

SAOUHSC_02169 chp chemotaxis inhibitory protein 1.7 4.4<br />

SAOUHSC_02161 eap MHC class II analog protein 1.8 11.0<br />

SAOUHSC_01114 efb fibrinogen-binding protein 1.9 2.7<br />

SAOUHSC_00816 emp extracellular matrix and plasma binding protein, putative -1.4 2.9<br />

SAOUHSC_00715 saeR response regulator, putative -1.2 2.1<br />

SAOUHSC_01115 hypothetical protein SAOUHSC_01115 1.8 3.0<br />

SAOUHSC_00354 hypothetical protein SAOUHSC_00354 -1.7 2.9<br />

SAOUHSC_00192 coa coagulase 6.3 1.8<br />

SAOUHSC_00400 hypothetical protein SAOUHSC_00400 9.1 1.7<br />

SAOUHSC_00399 hypothetical protein SAOUHSC_00399 10.4 1.9<br />

SAOUHSC_00818 nuc thermonuclease precursor -4.4 1.2<br />

SAOUHSC_02229 anti repressor 1.4 -2.9<br />

SAOUHSC_00182 hypothetical protein SAOUHSC_00182 -2.6 -2.8<br />

a Differentially regulated genes (significant with p* < 0.05 and with a minimal absolute fold change <strong>of</strong> 2) are indicated in bold.<br />

Differential expression <strong>of</strong> virulence associated genes<br />

Very important for colonization <strong>of</strong> the <strong>host</strong>, but also for internalization into <strong>host</strong> cells are the<br />

staphylococcal membrane bound adhesins, MSCRAMMs. Genes coding for fibrinogen binding<br />

protein A and B (fnbA, fnbB) and for clumping factor A and B (clfA, clfB) were induced in<br />

internalized staphylococci 2.5 h after start <strong>of</strong> infection. Differential expression <strong>of</strong> fnbA and clfA<br />

was still detectable at the 6.5 h time point whereas that <strong>of</strong> fnbB and clfB was not significant<br />

anymore, although a trend <strong>of</strong> induction was visible according to the fold change values. The<br />

control samples exhibited only less consistent expression changes. An exception might be the<br />

expression <strong>of</strong> fnbA and clfA in the 6.5 h serum/CO 2 control. Here, a trend <strong>of</strong> induction was visible,<br />

but the difference could not be determined statistically because <strong>of</strong> the number <strong>of</strong> replicates<br />

(Fig. R.5.14 A, B). Soluble adhesins bind to <strong>host</strong> structures. They can mediate bacterial cell<br />

attachment indirectly <strong>by</strong> involving linker molecules. Some alternatively exist in cell-surface<br />

associated forms. IsaB has an intermediate position as membrane bound and secreted protein,<br />

which was recently discovered to bind extracellular dsDNA and which is implicated in virulence,<br />

but whose exact function is still not identified (Mackey-Lawrence et al. 2009). In staphylococci<br />

internalized <strong>by</strong> S9 cells, expression <strong>of</strong> isaB, eap, coa, vwb, emp, and efb was induced in at least<br />

one <strong>of</strong> the two analyzed time points <strong>of</strong> internalization. In this group, the repression <strong>of</strong> coa and<br />

vwb occurred in the 2.5 h serum/CO 2 and 2.5 h anaerobic control samples, and vwb and efb were<br />

significantly less expressed in the exponential growth samples with which inoculation <strong>of</strong> S9 cell<br />

culture plates took place. Contrarily, the gene eap was two-fold induced in the 2.5 h serum/CO 2<br />

controls, and isaB showed induction in anaerobiosis (Fig. R.5.14 A, B, C).<br />

149


mean normalized intensity<br />

mean normalized intensity<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

A<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

fnbA fnbB clfA clfB isaB eap coa vwb emp efb<br />

exponential growth phase -1.8 -2.8 1.6 -1.7 1.3 -2.0 -2.2 -5.8 -1.6 -2.8<br />

2.5 h internalized 6.2 7.0 5.3 2.1 2.1 1.8 6.3 4.9 -1.4 1.9<br />

6.5 h internalized 6.4 3.1 2.8 2.1 -1.3 11.0 1.8 2.7 2.9 2.7<br />

2.5 h serum/CO 2 control 1.1 -2.9 2.5 -1.6 1.5 2.0 -3.6 -7.7 -1.9 -1.3<br />

6.5 h serum/CO 2 control 3.4 -2.0 9.2 -3.3 -2.1 4.7 -5.7 -4.2 -1.6 -1.3<br />

2.5 h anaerobic incubation 1.3 -1.2 3.3 -1.4 4.7 3.1 -2.7 -3.3 1.6 1.2<br />

B<br />

C<br />

10.0<br />

10.0<br />

fnbA<br />

eap<br />

fnbB<br />

coa<br />

clfA<br />

vwb<br />

1.0<br />

clfB<br />

isaB<br />

1.0<br />

emp<br />

efb<br />

0.1<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

0.1<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

Fig. R.5.14: Adhesins (MSCRAMMs and SERAMs) gene expression.<br />

A. Overview on fold change values relative to the baseline sample “1 h serum/CO 2 control” and on significance in statistical group<br />

comparisons. Differentially expressed genes are marked <strong>by</strong> gray filling for the corresponding sample condition/time point.<br />

Genes which were significant in statistical testing (p* < 0.05) but did not pass the absolute fold change cut<strong>of</strong>f 2 are indicated in bold<br />

without filling. The sample “6.5 h serum/CO 2 control” could not be included in statistical testing because <strong>of</strong> small group size (n = 2).<br />

B, C. Overview on mean normalized gene expression intensity for MSCRAMM (B) and SERAM (C) genes. After the global normalization<br />

<strong>by</strong> inter-chip scaling and detrending, each individual gene has been normalized to the expression level <strong>of</strong> the baseline sample “1 h<br />

serum/CO 2 control”. Although not being continuous data, values were depicted as line graphs instead <strong>of</strong> bar charts for better facility<br />

<strong>of</strong> inspection.<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2 control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h internalization;<br />

2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h co. − 6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic incubation)<br />

Toxins, especially lytic toxins, are one <strong>of</strong> the most outstanding groups <strong>of</strong> virulence factors<br />

because they directly harm the <strong>host</strong> cells. In the group <strong>of</strong> toxins, two bicomponent toxin gene<br />

pairs were observed to be induced in internalized staphylococci: lukD/lukE and hlgB/hlgC.<br />

Furthermore, alpha-hemolysin (hla) was induced. Induction for all genes was significant at the<br />

6.5 h time point except for hlgB, which was not significant although its fold change was the<br />

second highest in this group. The pair hlgB/hlgC was additionally already induced in internalized<br />

staphylococci 2.5 h after start <strong>of</strong> infection, and also in the 2.5 h serum/CO 2 control. Again, the<br />

6.5 h serum/CO 2 control exhibited an even further increased fold change, but could not be tested<br />

statistically (Fig. R.5.15 A, B).<br />

Further virulence factors, secreted and membrane bound enzymes, were observed with<br />

divergent expression pattern in internalized staphylococci. All six members A, B, C, D, E, F <strong>of</strong> the<br />

spl operon, coding for serine proteases with similarity to the V8 protease, were found to be<br />

induced at the 6.5 h time point <strong>of</strong> internalization. Four <strong>of</strong> them, A, B, C, and E, were additionally<br />

induced 2.5 h after start <strong>of</strong> infection. For this operon, a similar induction was observed in the<br />

serum/CO 2 control samples. Another exoenzyme, lipase (lip), was induced in internalized<br />

staphylococci, but the peak <strong>of</strong> induction occurred 2.5 h after start <strong>of</strong> infection, i. e. earlier than<br />

for the spl operon. In serum/CO 2 control, lip gene expression did not change after 2.5 h, but a<br />

strong induction was detectable after 6.5 h although statistical testing was not possible<br />

(Fig. R.5.16 A, B). A second group <strong>of</strong> enzymes exhibited repression in internalized staphylococci:<br />

hysA (hyaluronate lyase), htrA (serine protease, a membrane bound enzyme), nuc (nuclease), and<br />

sspA (glutamyl endopeptidase, V8 peptidase).<br />

150


mean normalized intensity<br />

mean normalized intensity<br />

mean normalized intensity<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

A<br />

B<br />

100.0<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

lukD lukE hlgB hlgC hla<br />

exponential growth phase -1.1 -1.9 -2.9 -5.6 -5.2<br />

2.5 h internalized -1.3 -1.3 3.4 5.3 1.9<br />

6.5 h internalized 6.5 9.0 13.4 17.9 3.2<br />

2.5 h serum/CO 2 control 1.1 1.0 3.1 3.5 1.3<br />

6.5 h serum/CO 2 control 1.4 1.1 8.0 8.9 1.7<br />

2.5 h anaerobic incubation -1.5 -2.0 3.4 5.0 2.4<br />

Fig. R.5.15: Staphylococcal toxin gene expression.<br />

A. Overview on fold change values relative to the baseline sample “1 h serum/CO 2 control” and on significance in statistical group<br />

comparisons. Differentially expressed genes are marked <strong>by</strong> gray filling for the corresponding sample condition/time point.<br />

Genes which were significant in statistical testing (p* < 0.05) but did not pass the absolute fold change cut<strong>of</strong>f 2 are indicated in bold<br />

without filling. The sample “6.5 h serum/CO 2 control” could not be included in statistical testing because <strong>of</strong> small group size (n = 2).<br />

B. Overview on mean normalized gene expression intensity. After the global normalization <strong>by</strong> inter-chip scaling and detrending, each<br />

individual gene has been normalized to the expression level <strong>of</strong> the baseline sample “1 h serum/CO 2 control”. Although not being<br />

continuous data, values were depicted as line graphs instead <strong>of</strong> bar charts for better facility <strong>of</strong> inspection.<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2 control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h internalization;<br />

2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h co. − 6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic incubation)<br />

10.0<br />

1.0<br />

0.1<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

lukD<br />

lukE<br />

hlgB<br />

hlgC<br />

hla<br />

A<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

splF splE splD splC splB splA lip hysA htrA nuc sspA sspB sspC<br />

exponential growth phase -1.3 -1.7 -1.1 -3.3 -5.4 -4.7 -1.4 -1.5 -1.2 -2.6 -1.4 -1.3 1.0<br />

2.5 h internalized 1.6 3.1 3.2 4.8 3.3 3.2 34.2 -2.5 -2.3 -4.4 -2.5 -1.4 -1.1<br />

6.5 h internalized 14.9 25.9 32.0 43.1 35.5 30.4 10.7 -3.0 -2.0 1.2 -1.2 -1.1 1.3<br />

2.5 h serum/CO 2 control 1.9 3.1 3.1 3.8 2.7 2.9 1.3 -2.6 -2.0 4.0 -1.3 -1.0 1.5<br />

6.5 h serum/CO 2 control 9.6 15.8 17.1 22.2 15.4 10.8 27.7 -3.4 -3.1 -1.0 -1.0 1.4 1.7<br />

2.5 h anaerobic incubation -1.0 1.5 1.5 2.2 1.5 1.7 1.9 -1.7 -1.5 1.0 1.1 -1.0 1.3<br />

B<br />

100.0<br />

C<br />

10.0<br />

splF<br />

hysA<br />

10.0<br />

1.0<br />

splE<br />

splD<br />

splC<br />

splB<br />

splA<br />

1.0<br />

htrA<br />

nuc<br />

sspA<br />

sspB<br />

sspC<br />

lip<br />

0.1<br />

0.1<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

Fig. R.5.16: Extracellular and membrane bound enzyme gene expression.<br />

A. Overview on fold change values relative to the baseline sample “1 h serum/CO 2 control” and on significance in statistical group<br />

comparisons. Differentially expressed genes are marked <strong>by</strong> gray filling for the corresponding sample condition/time point.<br />

Genes which were significant in statistical testing (p* < 0.05) but did not pass the absolute fold change cut<strong>of</strong>f 2 are indicated in bold<br />

without filling. The sample “6.5 h serum/CO 2 control” could not be included in statistical testing because <strong>of</strong> small group size (n = 2).<br />

B, C. Overview on mean normalized gene expression intensity for induced extracellular enzyme (B) and repressed extracellular and<br />

membrane bound enzyme (C) genes. After the global normalization <strong>by</strong> inter-chip scaling and detrending, each individual gene has<br />

been normalized to the expression level <strong>of</strong> the baseline sample “1 h serum/CO 2 control”. Although not being continuous data, values<br />

were depicted as line graphs instead <strong>of</strong> bar charts for better facility <strong>of</strong> inspection.<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2 control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h internalization;<br />

2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h co. − 6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic incubation)<br />

151


mean normalized intensity<br />

mean normalized intensity<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

Additionally, repression was observed for hysA, htrA, and nuc in 2.5 h serum/CO 2 control.<br />

While hysA and htrA were repressed in both analyzed time points <strong>of</strong> internalization, the<br />

repression <strong>of</strong> sspA was only detected 2.5 h after start <strong>of</strong> infection. The other two genes <strong>of</strong> the ssp<br />

operon were not detected as differentially expressed (Fig. R.5.16 A, C).<br />

Staphylococci secrete proteins which help the bacterium to evade the <strong>host</strong>’s immune<br />

response. In the experimental setting <strong>of</strong> the in vitro S9 infection and internalization model, some<br />

<strong>of</strong> them were induced in S. aureus RN1HG after internalization in S9 cells, especially at the 6.5 h<br />

time point: chp (chemotaxis inhibitory protein CHIPS), eap (extracellular adherence protein, also<br />

called MHC class II analog protein Map), and efb (extracellular fibrinogen-binding protein).<br />

Staphylokinase (sak) was repressed in internalized staphylococci at the 2.5 h time point<br />

(Fig. R.5.17 A, B). Staphylococci own two genes coding for superoxide dismutases, sodA and<br />

sodM. These two gene exhibited divergent regulation during internalization: sodA was induced<br />

and sodM was repressed in at least one <strong>of</strong> the two analyzed time points (Fig. R.5.17 A, C).<br />

A<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

chp eap efb sak sodA sodM<br />

exponential growth phase -7.9 -2.0 -2.8 -2.8 2.4 -1.2<br />

2.5 h internalized 1.7 1.8 1.9 -3.8 2.4 -1.7<br />

6.5 h internalized 4.4 11.0 2.7 -1.9 3.2 -3.0<br />

2.5 h serum/CO 2 control -1.2 2.0 -1.3 2.2 1.9 -1.1<br />

6.5 h serum/CO 2 control -3.1 4.7 -1.3 -2.7 2.7 -2.5<br />

2.5 h anaerobic incubation -4.6 3.1 1.2 1.1 1.3 -1.6<br />

B<br />

100.0<br />

C<br />

10.0<br />

10.0<br />

1.0<br />

chp<br />

eap<br />

efb<br />

sak<br />

1.0<br />

sodA<br />

sodM<br />

0.1<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

0.1<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

Fig. R.5.17: Immune evasion gene expression.<br />

A. Overview on fold change values relative to the baseline sample “1 h serum/CO 2 control” and on significance in statistical group<br />

comparisons. Differentially expressed genes are marked <strong>by</strong> gray filling for the corresponding sample condition/time point.<br />

Genes which were significant in statistical testing (p* < 0.05) but did not pass the absolute fold change cut<strong>of</strong>f 2 are indicated in bold<br />

without filling. The sample “6.5 h serum/CO 2 control” could not be included in statistical testing because <strong>of</strong> small group size (n = 2).<br />

B, C. Overview on mean normalized gene expression intensity for different immune evasion (B) and superoxide dismutase (C) genes.<br />

After the global normalization <strong>by</strong> inter-chip scaling and detrending, each individual gene has been normalized to the expression level<br />

<strong>of</strong> the baseline sample “1 h serum/CO 2 control”. Although not being continuous data, values were depicted as line graphs instead <strong>of</strong><br />

bar charts for better facility <strong>of</strong> inspection.<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2 control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h internalization;<br />

2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h co. − 6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic incubation)<br />

Changes in the cell wall allow staphylococci defense against and evasion <strong>of</strong> the immune<br />

response. The dlt operon for example is responsible for the incorporation <strong>of</strong> D-alanine into the<br />

teichoic acids, which leads to a reduction <strong>of</strong> negative charge <strong>of</strong> the cell wall. Thus, these bacterial<br />

cells are less vulnerable <strong>by</strong> antimicrobial cationic peptides due to a reduced attraction.<br />

Surprinsingly, this operon was temporarily repressed at the 2.5 h time point <strong>of</strong> internalization. At<br />

the same time and also 6.5 h after start <strong>of</strong> infection, the genes <strong>of</strong> cell wall modulating enzymes<br />

lytM (peptidoglycan hydrolase, endopeptidase) and ssaA (secretory antigen precursor, amidase)<br />

were induced in internalized staphylococci (Fig. R.5.18 A, B).<br />

152


mean normalized intensity<br />

mean normalized intensity<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

A<br />

B<br />

10.0 10,0<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

dltA dltB dltC dltD lytM ssaA<br />

exponential growth phase 1.0 -1.0 -1.0 -1.2 2.6 1.2<br />

2.5 h internalized -2.5 -2.2 -1.6 -2.2 2.4 3.0<br />

6.5 h internalized -1.5 -1.3 -1.2 -1.3 2.1 2.9<br />

2.5 h serum/CO 2 control 1.0 -1.0 1.1 -1.1 1.4 1.3<br />

6.5 h serum/CO 2 control -2.5 -2.4 -2.5 -2.4 1.5 -1.9<br />

2.5 h anaerobic incubation -2.0 -2.2 -1.9 -2.7 -3.3 -2.5<br />

1.0 1,0<br />

dltA<br />

dltB<br />

dltC<br />

dltD<br />

lytM<br />

ssaA<br />

Fig. R.5.18: Cell wall associated gene expression.<br />

A. Overview on fold change values relative to the baseline sample “1 h serum/CO 2 control” and on significance in statistical group<br />

comparisons. Differentially expressed genes are marked <strong>by</strong> gray filling for the corresponding sample condition/time point.<br />

Genes which were significant in statistical testing (p* < 0.05) but did not pass the absolute fold change cut<strong>of</strong>f 2 are indicated in bold<br />

without filling. The sample “6.5 h serum/CO 2 control” could not be included in statistical testing because <strong>of</strong> small group size (n = 2).<br />

B. Overview on mean normalized gene expression intensity. After the global normalization <strong>by</strong> inter-chip scaling and detrending, each<br />

individual gene has been normalized to the expression level <strong>of</strong> the baseline sample “1 h serum/CO 2 control”. Although not being<br />

continuous data, values were depicted as line graphs instead <strong>of</strong> bar charts for better facility <strong>of</strong> inspection.<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2 control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h internalization;<br />

2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h co. − 6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic incubation)<br />

0.1 0,1<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

Bi<strong>of</strong>ilm formation is another option for S. aureus to protect the bacterial cells from the<br />

immune system and antimicrobials. Here, bacterial cells are surrounded <strong>by</strong> a polysaccharide<br />

matrix, composed <strong>of</strong> poly-N-acetylglucosamine (PNAG). The ica operon encodes the proteins<br />

necessary for PNAG biosynthesis. Of this operon, the icaB gene was repressed in internalized and<br />

control samples <strong>of</strong> the infection experiment. The icaC and icaD genes behaved similarly, but were<br />

significantly repressed only at the 2.5 h time point <strong>of</strong> internalization. The fourth gene <strong>of</strong> the<br />

operon, icaA, and the separately encoded repressor gene, icaR, were not differentially expressed<br />

(Fig. R.5.19 A, B). Furthermore, it was obvious that the expression intensity <strong>of</strong> icaADBC was low.<br />

A<br />

B<br />

10,0 10.0<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

icaR icaA icaD icaB icaC<br />

exponential growth phase -1.0 2.0 1.6 1.7 1.6<br />

2.5 h internalized -1.2 1.4 -2.2 -3.7 -2.8<br />

6.5 h internalized -1.0 1.0 -1.3 -2.2 -1.5<br />

2.5 h serum/CO 2 control -1.4 1.3 1.1 -2.2 1.1<br />

6.5 h serum/CO 2 control -1.2 1.2 1.5 -1.9 -1.4<br />

2.5 h anaerobic incubation -1.3 1.1 -1.6 -2.4 -1.4<br />

Fig. R.5.19: Bi<strong>of</strong>ilm associated gene expression.<br />

A. Overview on fold change values relative to the baseline sample “1 h serum/CO 2 control” and on significance in statistical group<br />

comparisons. Differentially expressed genes are marked <strong>by</strong> gray filling for the corresponding sample condition/time point.<br />

Genes which were significant in statistical testing (p* < 0.05) but did not pass the absolute fold change cut<strong>of</strong>f 2 are indicated in bold<br />

without filling. The sample “6.5 h serum/CO 2 control” could not be included in statistical testing because <strong>of</strong> small group size (n = 2).<br />

B. Overview on mean normalized gene expression intensity. After the global normalization <strong>by</strong> inter-chip scaling and detrending, each<br />

individual gene has been normalized to the expression level <strong>of</strong> the baseline sample “1 h serum/CO 2 control”. Although not being<br />

continuous data, values were depicted as line graphs instead <strong>of</strong> bar charts for better facility <strong>of</strong> inspection.<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2 control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h internalization;<br />

2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h co. − 6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic incubation)<br />

1,0 1.0<br />

0,1 0.1<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

icaR<br />

icaA<br />

icaD<br />

icaB<br />

icaC<br />

153


mean normalized intensity<br />

mean normalized intensity<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

Besides the beforementioned two-component system SaeRS, whose gene expression was<br />

increased, further regulators were differentially expressed in the internalization experiment. The<br />

transcription factors SarT, SarU, and Rot belong to the virulence associated Sar family. SarT and<br />

sarU were repressed in internalized staphylococci at the 6.5 h time point, whereas rot was<br />

repressed at the 2.5 h time point (Fig. R.5.20 A, B). Expression intensity <strong>of</strong> sarU gene was very<br />

low.<br />

A<br />

B<br />

10.00<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

saeR saeS sarT sarU rot<br />

exponential growth phase -1.8 -1.8 2.1 -1.0 -1.6<br />

2.5 h internalized -1.2 -1.2 -1.6 -1.7 -2.4<br />

6.5 h internalized 2.1 1.7 -4.6 -2.2 -1.7<br />

2.5 h serum/CO 2 control 1.7 1.5 -4.5 2.6 1.6<br />

6.5 h serum/CO 2 control 2.5 1.9 -13.3 1.2 -1.2<br />

2.5 h anaerobic incubation 1.9 1.5 -1.0 1.1 1.1<br />

Fig. R.5.20: Expression <strong>of</strong> regulator genes.<br />

A. Overview on fold change values relative to the baseline sample “1 h serum/CO 2 control” and on significance in statistical group<br />

comparisons. Differentially expressed genes are marked <strong>by</strong> gray filling for the corresponding sample condition/time point.<br />

Genes which were significant in statistical testing (p* < 0.05) but did not pass the absolute fold change cut<strong>of</strong>f 2 are indicated in bold<br />

without filling. The sample “6.5 h serum/CO 2 control” could not be included in statistical testing because <strong>of</strong> small group size (n = 2).<br />

B. Overview on mean normalized gene expression intensity. After the global normalization <strong>by</strong> inter-chip scaling and detrending, each<br />

individual gene has been normalized to the expression level <strong>of</strong> the baseline sample “1 h serum/CO 2 control”. Although not being<br />

continuous data, values were depicted as line graphs instead <strong>of</strong> bar charts for better facility <strong>of</strong> inspection.<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2 control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h internalization;<br />

2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h co. − 6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic incubation)<br />

1.00<br />

0.10<br />

0.01<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

saeR<br />

saeS<br />

sarT<br />

sarU<br />

rot<br />

A<br />

B<br />

10.0<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

prsA vraR vraS<br />

exponential growth phase 1.1 -1.1 -1.1<br />

2.5 h internalized 2.0 1.8 1.6<br />

6.5 h internalized 1.6 1.4 1.3<br />

2.5 h serum/CO 2 control -1.2 -1.2 -1.1<br />

6.5 h serum/CO 2 control -1.9 -1.1 -1.0<br />

2.5 h anaerobic incubation 1.9 1.4 1.6<br />

1.0<br />

0.1<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

prsA<br />

vraR<br />

vraS<br />

Fig. R.5.21: Gene expression <strong>of</strong> prsA and its regulatory two-component system’s genes vraR and vraS.<br />

A. Overview on fold change values relative to the baseline sample “1 h serum/CO 2 control” and on significance in statistical group<br />

comparisons. Differentially expressed genes are marked <strong>by</strong> gray filling for the corresponding sample condition/time point.<br />

Genes which were significant in statistical testing (p* < 0.05) but did not pass the absolute fold change cut<strong>of</strong>f 2 are indicated in bold<br />

without filling. The sample “6.5 h serum/CO 2 control” could not be included in statistical testing because <strong>of</strong> small group size (n = 2).<br />

B. Overview on mean normalized gene expression intensity. After the global normalization <strong>by</strong> inter-chip scaling and detrending, each<br />

individual gene has been normalized to the expression level <strong>of</strong> the baseline sample “1 h serum/CO 2 control”. Although not being<br />

continuous data, values were depicted as line graphs instead <strong>of</strong> bar charts for better facility <strong>of</strong> inspection.<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2 control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h internalization;<br />

2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h co. − 6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic incubation)<br />

154


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

Proteome analysis <strong>of</strong> internalized staphylococci, which was performed <strong>by</strong> Sandra Scharf,<br />

revealed the increased abundance <strong>of</strong> PrsA. In parallel, the response regulator VraR <strong>of</strong> the twocomponent<br />

system VraRS was up-regulated. Also transcriptome analysis revealed induction <strong>of</strong><br />

prsA in internalized staphylococci at the time point 2.5 h after start <strong>of</strong> infection. The regulatory<br />

genes vraR and vraS were not differentially expressed. Nevertheless, with a fold change <strong>of</strong> 1.8<br />

and 1.6 for vraR and vraS, respectively, a tendency for an induction <strong>of</strong> gene expression was<br />

visible, which fitted well to the proteome data (Fig. R.5.21 A, B).<br />

The VraRS regulon contains 46 genes including vraR and vraS (Kuroda et al. 2003). Of the<br />

published 46 S. aureus N315 LocusTags, 45 could be mapped to S. aureus NCTC8325 LocusTags<br />

(NCBI’s EntrezGenome Gene Plot; http://www.ncbi.nlm.nih.gov/sutils/geneplot.cgi). When asking<br />

for the fraction <strong>of</strong> the VraRS regulon which was regulated in internalized S. aureus RN1HG in S9<br />

cells, 22 % (2.5 h after start <strong>of</strong> infection) and 11 % (6.5 h after start <strong>of</strong> infection) were differentially<br />

expressed upon internalization. In detail, 7 genes were induced 2.5 h after start <strong>of</strong> infection, <strong>of</strong><br />

which 4 were still induced at the later time point <strong>of</strong> 6.5 h after start <strong>of</strong> infection. Repressed gene<br />

expression was also observed: Repression was detected for 3 genes at the early 2.5 h time point,<br />

and one <strong>of</strong> these genes still was repressed at the later 6.5 h time point (Fig. R.5.22 A, B).<br />

A<br />

VraRS regulon according to<br />

Kuroda et al. 2003<br />

B<br />

VraRS regulon according to<br />

Kuroda et al. 2003<br />

35<br />

40<br />

7 3<br />

0<br />

395 360<br />

0<br />

4 1<br />

0<br />

315 307<br />

0<br />

induced sequences in<br />

“2.5 h internalization” vs.<br />

“1 h serum/CO 2 control”<br />

repressed sequences in<br />

“2.5 h internalization” vs.<br />

“1 h serum/CO 2 control”<br />

induced sequences in<br />

“6.5 h internalization” vs.<br />

“1 h serum/CO 2 control”<br />

repressed sequences in<br />

“6.5 h internalization” vs.<br />

“1 h serum/CO 2 control”<br />

Fig. R.5.22: Differential expression <strong>of</strong> genes belonging to the VraRS regulon in internalized S. aureus RN1HG in S9 cells.<br />

Genes known to be regulated <strong>by</strong> VraRS (according to Kuroda et al. 2003) were compared to S. aureus RN1HG S9 internalization gene<br />

expression signatures separately for induced and repressed genes.<br />

A. 2.5 h after start <strong>of</strong> infection.<br />

B. 6.5 h after start <strong>of</strong> infection.<br />

Differentially regulated metabolic genes<br />

First, BIOCYC “omics-Viewer” pathway mapping (BIOCYC, SRI International, CA, USA,<br />

http://biocyc.org/expression.html) was applied for a comprehensive overview on changes in<br />

gene expression <strong>of</strong> metabolic enzymes in internalized and control staphylococci. This tool allows<br />

the display <strong>of</strong> gene expression data on highly abstracted metabolic pathway schemes and<br />

therefore an intuitive comprehension <strong>of</strong> processes in the selected experimental setup.<br />

In comparison to the baseline <strong>of</strong> 2.5 h serum/CO 2 control samples, differentially expressed<br />

sequences in a) 2.5 h internalization, b) 6.5 h internalization, c) 2.5 h serum/CO 2 control, and<br />

d) 2.5 h anaerobic incubation were included in the analysis.<br />

Several changes in the expression <strong>of</strong> metabolic genes became visible, but certain functionally<br />

associated groups <strong>of</strong> genes accumulated changes in expression (Fig. R.5.23).<br />

155


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

A<br />

phosphonate/<br />

phosphate<br />

transporters<br />

uracil<br />

permease<br />

oligopeptide<br />

transporters<br />

spermidine/<br />

putrescine<br />

transporter<br />

purine and<br />

pyrimidine<br />

nucleotide<br />

biosynthesis<br />

sodium/sulfate, fructose, glycine<br />

betaine, choline, L-lactate, xanthine,<br />

urea transporter/symporter/permease<br />

acetoin<br />

biosynthesis<br />

different<br />

fermentation<br />

pathways<br />

degradation <strong>of</strong> glycerol, mannitol,<br />

galactose, fructose, nucleosides,<br />

citrulline/ornithine, arg, his, ala<br />

and others<br />

maltose transporter<br />

protease<br />

proline betaine transporter<br />

amino acid transporter<br />

phospho-transferase system<br />

(PTS); glucose-and mannitolspecific<br />

components<br />

glycolysis<br />

fatty acid<br />

biosynthesis<br />

arginine/<br />

ornithine<br />

antiporter<br />

amino acid<br />

biosynthesis<br />

sodium-dependent tr.,<br />

phosphate transporter<br />

sodium/glutamate symporter,<br />

proline uptake protein,<br />

citrate transporter<br />

tRNA charging<br />

pathways<br />

anaerobic<br />

respiration<br />

aerobic<br />

respiration<br />

gluconate<br />

permease<br />

protein modification: oxoacylreductase,<br />

ketoacyl-reductase,<br />

oxoacyl-synthase reactions at<br />

different C positions<br />

B<br />

choline, L-lactate transporter<br />

different<br />

fermentation<br />

pathways<br />

acetoin<br />

biosynthesis<br />

degradation <strong>of</strong> glycerol, mannitol,<br />

galactose, fructose, nucleosides,<br />

citrulline/ornithine, arg, his, ala<br />

and others<br />

protease<br />

phosponate/<br />

phosphate and<br />

oligopeptide<br />

transporters<br />

spermidine/<br />

putrescine<br />

transporter<br />

proline betaine transporter<br />

amino acid transporter<br />

phospho-transferase system<br />

(PTS); glucose-and mannitolspecific<br />

components<br />

purine and<br />

pyrimidine<br />

nucleotide<br />

biosynthesis<br />

glycolysis<br />

fatty acid<br />

biosynthesis<br />

arginine/<br />

ornithine<br />

antiporter<br />

amino acid<br />

biosynthesis<br />

phosphate<br />

transporter<br />

tRNA charging<br />

sodium/glutamate symporter pathways<br />

proline uptake protein<br />

anaerobic<br />

respiration<br />

aerobic<br />

respiration<br />

gluconate<br />

permease<br />

protein modification: oxoacylreductase,<br />

ketoacyl-reductase,<br />

oxoacyl-synthase reactions at<br />

different C positions<br />

156


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

C<br />

peptidoglycan<br />

biosynthesis<br />

fructose, glycine betaine<br />

permease/transporter<br />

different<br />

fermentation<br />

pathways<br />

acetoin<br />

biosynthesis<br />

degradation <strong>of</strong> glycerol, mannitol,<br />

galactose, fructose, nucleosides,<br />

citrulline/ornithine, arg, his, ala<br />

and others<br />

ammonium transporter<br />

gluconeogenesis<br />

uracil<br />

permease<br />

oligopeptide<br />

transporters<br />

spermidine/<br />

putrescine<br />

transporter<br />

protease<br />

proline betaine transporter<br />

branched and standard<br />

amino acid transporter<br />

phospho-transferase system<br />

(PTS); glucose-and mannitolspecific<br />

components<br />

purine and<br />

pyrimidine<br />

nucleotide<br />

biosynthesis<br />

glycolysis<br />

fatty acid<br />

biosynthesis<br />

amino acid<br />

biosynthesis<br />

sodium<br />

glutamate<br />

symporter<br />

proline uptake protein<br />

tRNA charging<br />

pathways<br />

anaerobic<br />

respiration<br />

aerobic<br />

respiration<br />

protein modification: oxoacylreductase,<br />

ketoacyl-reductase,<br />

oxoacyl-synthase reactions at<br />

different C positions<br />

D<br />

fructose, glycine betaine, choline,<br />

transporter/permease<br />

Na + /H + antiporter<br />

Mo 2+ transporter<br />

different<br />

fermentation<br />

pathways<br />

acetoin<br />

biosynthesis<br />

degradation <strong>of</strong> glycerol, mannitol,<br />

galactose, fructose, nucleosides,<br />

citrulline/ornithine, arg, his, ala<br />

and others<br />

S-, L-lactate permease<br />

proline betaine transporter<br />

monovalent<br />

cation/proton<br />

antiporter<br />

oligopeptide<br />

transporter<br />

ferrichrome<br />

transport<br />

permease;<br />

monovalent<br />

cation/proton<br />

antiporter;<br />

spermidine/<br />

putrescine<br />

transporter<br />

fatty acid<br />

biosynthesis<br />

arginine/<br />

ornithine<br />

antiporter<br />

amino acid<br />

biosynthesis<br />

nickel permease,<br />

Na + /H + antiporter<br />

sodium/glutamate symporter<br />

proline uptake protein<br />

tRNA charging<br />

pathways<br />

Na + /H + antiporter<br />

ferrichrome ABC<br />

transporter<br />

branched amino acid<br />

transporter<br />

standard amino acid<br />

transporter<br />

phospho-transferase system<br />

(PTS); glucose-and mannitolspecific<br />

components<br />

protein modification: oxoacylreductase,<br />

ketoacyl-reductase,<br />

oxoacyl-synthase reactions at<br />

different C positions<br />

Fig. R.5.23: The influence <strong>of</strong> internalization and control treatment on gene expression in staphylococcal NCTC8325 metabolic<br />

pathways (modified from omics-viewer(s) <strong>of</strong> BIOCYC, SRI International, CA, USA, http://biocyc.org/expression.html).<br />

Nodes represent metabolites, and lines indicate reactions. The metabolic reactions are colored according to the enzymes’ gene<br />

expression regulation in the samples <strong>of</strong> 2.5 h internalization (A), 6.5 h internalization (B), 2.5 h serum/CO 2 control (C), and 2.5 h<br />

anaerobic incubation; each sample was compared versus the baseline <strong>of</strong> 1 h serum/CO 2 control. Red marks an increase and yellow a<br />

decrease <strong>of</strong> expression in the sample. The display is limited to genes which were significant with p* < 0.05 in statistical testing and<br />

whose mean absolute fold change exceeded 2.<br />

157


mean normalized intensity<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

At the early internalization time point, several genes <strong>of</strong> purine and pyrimidine nucleotide<br />

biosynthesis pathways were repressed. Amino acid biosynthesis genes exhibited increase <strong>of</strong><br />

expression in both analyzed internalization time points. Furthermore, genes coding for<br />

respiration chain components were induced in expression in internalized staphylococci. In<br />

internalized staphylococci, several tRNA synthetase genes and some glycolytic enzyme genes<br />

were observed to be repressed. Finally, transporters and permeases for sugar molecules,<br />

phosphate, oligopeptides, and other substances were differentially expressed (Fig. R.5.23 A, B).<br />

Serum/CO 2 control samples featured in some aspects similar changes in metabolic enzyme’s gene<br />

expression pattern, e. g. amino acid biosynthesis, glycolysis, tRNA synthetases. On the other<br />

hand, aerobic respiration, sugar uptake phosphotransferase system, and some enzymes <strong>of</strong><br />

peptidoglycan biosynthesis were repressed in this group. Anaerobic incubation instead did not<br />

provoke changes in expression <strong>of</strong> glycolytic enzymes or respiration chain components<br />

(Fig. R.5.23 C, D).<br />

Subsequently to the first approach <strong>of</strong> BIOCYC pathway mapping, selected genes were analyzed<br />

in more detail. Here, the example <strong>of</strong> purine biosynthesis was chosen. Although not all genes were<br />

significantly differentially expressed, the gene expression changes were highly similar in the<br />

complete pur operon and guaAB, which are separately encoded but are also involved in purine<br />

biosynthesis. Five genes (purA, purC, purE, purK, purS) were differentially expressed in both<br />

analyzed internalization time points, six genes (purD, purF, purL, purQ, guaA, guaB) were<br />

differentially expressed 2.5 h after start <strong>of</strong> infection. Three genes, purH, purM, and purN,<br />

exhibited fold change values between −2.6 and −2.8, but were not significant in statistical testing.<br />

Finally, purB was significant in the statistical test in both analyzed time points <strong>of</strong> internalization,<br />

but did not pass the minimal absolute fold change cut<strong>of</strong>f <strong>of</strong> 2. In the 2.5 h serum/CO 2 control,<br />

differential expression occurred only for purA and in trend for purB, whereas almost all genes<br />

exhibited fold change values around −3 to −4 in the 6.5 h serum/CO 2 control. Exceptions were<br />

guaA and guaB, whose fold change values were still around 1, and the repressor gene purR,<br />

which did not exhibit expression changes in any <strong>of</strong> the analyzed experimental conditions<br />

(Fig. R.5.24, Fig. R.5.25).<br />

A<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

purA purB purC purD purE purF purH purK purL purM purN purQ purR purS guaA guaB<br />

exponential growth phase -1.1 -1.0 1.2 1.3 1.2 1.1 1.2 1.2 1.1 1.1 1.2 1.1 -1.2 1.0 -1.0 -1.1<br />

2.5 h internalized -4.8 -1.6 -4.5 -2.6 -6.2 -3.0 -2.6 -5.4 -3.4 -2.8 -2.6 -4.0 1.1 -4.9 -2.0 -2.3<br />

6.5 h internalized -4.8 -1.6 -2.2 -1.1 -3.4 -1.2 -1.2 -2.9 -1.5 -1.2 -1.1 -1.9 1.3 -2.3 -1.3 -1.4<br />

2.5 h serum/CO 2 control -6.4 -1.6 1.0 1.1 1.0 -1.1 -1.0 1.1 -1.0 -1.0 1.0 1.0 1.3 -1.1 1.4 1.3<br />

6.5 h serum/CO 2 control -28.2 -2.7 -3.2 -3.4 -3.8 -4.1 -4.3 -3.1 -4.0 -3.9 -3.6 -4.0 1.1 -3.9 1.1 -1.0<br />

2.5 h anaerobic incubation -2.1 -1.7 1.4 1.1 1.7 1.3 1.2 1.5 1.3 1.3 1.2 1.5 -1.1 1.1 1.0 1.0<br />

Fig. R.5.24: Purine biosynthesis gene expression.<br />

A. Overview on fold change values relative to the baseline<br />

sample “1 h serum/CO 2 control” and on significance in statistical<br />

group comparisons. The sample “6.5 h serum/CO 2 control” could<br />

not be included in statistical testing because <strong>of</strong> small group size<br />

(n = 2).<br />

B. Overview on mean normalized gene expression intensity.<br />

After the global normalization <strong>by</strong> inter-chip scaling and<br />

detrending, each individual gene has been normalized to the<br />

expression level <strong>of</strong> the baseline sample “1 h serum/CO 2 control”<br />

(1 h co.). Although not being continuous data, values were<br />

depicted as line graphs instead <strong>of</strong> bar charts for better facility <strong>of</strong><br />

inspection.<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2<br />

control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h<br />

internalization; 2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h co. −<br />

6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic<br />

incubation)<br />

B<br />

1.00<br />

0.10<br />

0.01<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

purA<br />

purB<br />

purC<br />

purD<br />

purE<br />

purF<br />

purH<br />

purK<br />

purL<br />

purM<br />

purN<br />

purQ<br />

purR<br />

purS<br />

guaA<br />

guaB<br />

158


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

01014<br />

purF<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

01018<br />

purD<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

01016<br />

purN<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

01013<br />

purL<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

01012<br />

purQ<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

01011<br />

purS<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

01015<br />

purM<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

01009<br />

purK<br />

01008<br />

purE<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

01010<br />

purC<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

02126<br />

purB<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

00467<br />

purR<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

01017<br />

purH<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

00019<br />

purA<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

02126<br />

purB<br />

AMP<br />

PRPP<br />

I-5‘-<br />

GMP<br />

PRPP : 5-phosphoribosyl-1-pyrophosphate<br />

I-5‘- : inosine-5‘-phosphate<br />

AMP : adenosine-5‘-monophosphate<br />

GMP : guanosine-5‘-monophosphate<br />

00374<br />

guaB<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

00375<br />

guaA<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

-8<br />

Fig. R.5.25: Purine biosynthesis pathway and gene expression.<br />

Pathway reactions and gene LocusTags were extracted from omics-viewer(s) <strong>of</strong> BIOCYC (BIOCYC, SRI International, CA, USA,<br />

http://biocyc.org/expression.html). Arrows represent reactions and dots substrates/products when these are not specially named.<br />

Numbers next to enzyme reactions will result in the LocusTag <strong>of</strong> the gene when combined with “SAOUHSC_*” instead <strong>of</strong> the wildcard<br />

character. Bar charts indicate fold change values <strong>of</strong> internalized samples in reference to the baseline sample <strong>of</strong> 1 h serum/CO 2 control.<br />

Values for the 2.5 h internalized samples are depicted in light gray (first column) and those <strong>of</strong> the 6.5 h internalized samples in dark<br />

gray (second column).<br />

A third analysis <strong>of</strong> metabolic gene expression involved the KEGG pathway maps (KEGG: Kyoto<br />

Encyclopedia <strong>of</strong> Genes and Genomes). Already BIOCYC pathway mapping had revealed changes in<br />

amino acid biosynthesis and glycolysis. Corresponding pathways and additionally pathways <strong>of</strong><br />

TCA and urea cycle were selected in KEGG specific for S. aureus NCTC8325. The strain’s genes,<br />

which were mapped <strong>by</strong> KEGG to the pathways, were extracted and checked for differential<br />

expression in at least one time point after internalization in S9 cells. The biosynthesis pathways<br />

for asparagine, histidine, glutamine, serine, tryptophan, leucine, valine, isoleucine, lysine, and<br />

threonine were induced to a large extent (Fig. R.5.26). A fraction <strong>of</strong> associated genes were not<br />

included in the lists <strong>of</strong> differentially expressed genes. A further analysis <strong>of</strong> these genes identified<br />

some <strong>of</strong> them to be either significant in statistical testing without passing the minimal absolute<br />

fold change cut<strong>of</strong>f or to pass that cut<strong>of</strong>f without reaching statistical significance. These<br />

expression changes were rated as trend, which in many cases further confirmed the findings <strong>of</strong><br />

expression changes in the metabolic pathways. In addition to induced expression <strong>of</strong> amino acid<br />

biosynthesis genes, induction <strong>of</strong> four TCA cycle enzyme genes was observed, which was even<br />

more substantiated <strong>by</strong> a trend <strong>of</strong> increase for further four genes. Contrarily, especially genes<br />

encoding enzymes <strong>of</strong> irreversible glycolysis reactions were repressed. Finally, genes coding for<br />

enzymes <strong>of</strong> anabolic reactions during gluconeogenesis were induced in internalized<br />

staphylococci. Using energy, their gene products reverse the reactions whose equilibrium is<br />

normally set to the side <strong>of</strong> glucose degradation products.<br />

Repressed tRNA synthetase gene expression has already been mentioned before. The<br />

synthetase genes for tyrosine, leucine, threonine, phenylalanine, serine, aspartate, and histidine<br />

(tyrS, leuS, thrS, pheS, pheT, serS, aspS, hisS) were repressed in at least one time point <strong>of</strong><br />

internalization, many <strong>of</strong> them even in both (Fig. R.5.27 A). Further genes (ileS, alaS, metS, asnS,<br />

gltX, trpS) exhibited a trend <strong>of</strong> repression in at least one time point <strong>of</strong> internalization with<br />

p* < 0.05, but an absolute fold change <strong>of</strong> less than 2 (Fig. R.5.27 B). Three genes (valS, glyS, cysS)<br />

behaved contrarily and possessed a trend <strong>of</strong> induction in one time point <strong>of</strong> internalization<br />

(Fig. R.5.27 C). Finally, the remaining three tRNA synthetase genes (argS, proS, lysS) showed no<br />

change in expression (Fig. R.5.27 D).<br />

159


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

purine<br />

metabolism<br />

glycerol<br />

01276<br />

glpK<br />

lactate<br />

00206 ldh1<br />

02922 ldh2<br />

pyruvate<br />

01818 ald1 01452 ald2<br />

alanine<br />

asparagine<br />

01497<br />

ahsA<br />

aspartate<br />

00899<br />

argG<br />

imidazolacetolphosphate<br />

00733<br />

hisC<br />

02607<br />

hutU<br />

AICAR<br />

03008 hisF<br />

03010 hisG<br />

03009<br />

hisA<br />

alanine aspartate<br />

02916<br />

panD<br />

glycolysis /<br />

gluconeogenesis<br />

glucose<br />

01430 crr (PTSIIA)<br />

00209 (PTSIIBC)<br />

00900 pgi<br />

fru-6-P<br />

02822 fbp<br />

01807 pfkA<br />

fru-1,6-BP<br />

02366 fba<br />

DHAP<br />

00797<br />

tpiA<br />

GA-3-P<br />

01794 gap2<br />

00795 gap1<br />

00796 pgk<br />

01833<br />

serA<br />

00798 pgm<br />

01910 pckA<br />

PEP<br />

pyruvate<br />

01806<br />

pykA<br />

02289<br />

ilvA<br />

01064 pycA<br />

00898<br />

argH<br />

malate<br />

fumarate<br />

01983 fumC<br />

oxalo<br />

acetate<br />

01103 sdhC<br />

01104 sdhA<br />

01105 sdhB<br />

succinate<br />

01216<br />

sucC<br />

01218<br />

sucD<br />

01802<br />

citZ<br />

citrate<br />

succinyl-<br />

CoA<br />

TCA<br />

cycle<br />

03013<br />

hisD<br />

histidine<br />

02607<br />

hutU<br />

03014<br />

hisG<br />

phosphoribosylpyrophosphate<br />

pentose<br />

phosphate<br />

pathway<br />

01394<br />

lysC<br />

01395<br />

asd<br />

01396<br />

dapA<br />

01397<br />

dapB<br />

01398<br />

dapD<br />

01319<br />

thrA<br />

01322 thrB<br />

homoserine<br />

01320 dhoM<br />

00013<br />

acetylhomoserine<br />

01321 thrC<br />

cystathionine<br />

02286<br />

leuB<br />

threonine<br />

02289 ilvA<br />

2-oxobutanoate<br />

cysteine<br />

02287 leuC<br />

02288 leuD<br />

pyruvate<br />

02282<br />

ilvB<br />

acetyl-CoA<br />

indole<br />

acetaldehyde<br />

indole<br />

acetate<br />

00153<br />

ipdC<br />

indole<br />

pyruvate<br />

indole<br />

serine<br />

01371<br />

trpB<br />

tryptophan<br />

01371<br />

trpB<br />

01846<br />

acsA<br />

00132 aldA<br />

02363<br />

(02142 aldA2)<br />

01347<br />

citB<br />

01416<br />

sucB<br />

01418<br />

sucA<br />

01801 citC<br />

00895 gudB<br />

NH 3<br />

02606<br />

hutI<br />

02610<br />

hutG<br />

glutamate<br />

02869<br />

rocA<br />

00148<br />

arcJ<br />

2-oxoglutarate<br />

pyrroline-5-<br />

carboxylate<br />

02244<br />

01401<br />

lysA<br />

lysine<br />

homocysteine<br />

methionine<br />

isoleucine<br />

02284<br />

ilvC<br />

02281<br />

ilvD<br />

valine<br />

02285 leuA<br />

02287 leuC<br />

02288 leuD<br />

02919 panB<br />

02739<br />

panE<br />

pantoate<br />

00286<br />

leuB<br />

spontaneous<br />

01369<br />

trpC<br />

01370<br />

trpF<br />

chorismate<br />

prephenate<br />

01364<br />

tyrA<br />

00113 adhE<br />

00608 adhA<br />

00733 hisC<br />

phenylalanine tyrosine<br />

argininosuccinate<br />

car<strong>by</strong>moylphosphate<br />

citruline<br />

02968<br />

arcB<br />

02969<br />

arcA<br />

00898<br />

argH<br />

00149<br />

argC<br />

00150<br />

urea<br />

cycle<br />

00894<br />

proline<br />

02409 arg, rocF<br />

urea<br />

glutamate<br />

semialdehyde<br />

leucine<br />

acetyl-CoA acetate acetaldehyde ethanol<br />

ornithine arginine<br />

Fig. R.5.26: Pathway map <strong>of</strong> amino acid biosyntheses, glycolysis/gluconeogenesis, TCA cycle and urea cycle.<br />

Pathway reactions were extracted from KEGG (KEGG: Kyoto Encyclopedia <strong>of</strong> Genes and Genomes, http://www.genome.jp/kegg/) and<br />

arranged according to genes differentially expressed in S9 cell internalized staphylococci in reference to the baseline sample <strong>of</strong> 1 h<br />

serum/CO 2 control. Arrows represent reactions and dots substrates/products when these are not specially named. Numbers next to<br />

enzyme reactions will result in the LocusTag <strong>of</strong> the gene when combined with “SAOUHSC_*” instead <strong>of</strong> the wildcard character.<br />

Colored arrows in combination with colored gene names indicate differential expression, whereas a colored gene name alone (with a<br />

black arrow) indicates a trend <strong>of</strong> regulation when only one criterium, p* < 0.05 or minimal absolute fold change > 2, is fulfilled.<br />

Induction <strong>of</strong> gene expression is marked in red and repression in green color.<br />

160


mean normalized intensity<br />

mean normalized intensity<br />

mean normalized intensity<br />

mean normalized intensity<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

A<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

tyrS leuS thrS pheS pheT serS aspS hisS ileS alaS metS asnS gltX trpS valS glyS cysS argS proS lysS<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

exponential growth phase -1.1 -1.0 -1.4 -1.2 -1.3 -1.2 -1.3 -1.2 -1.2 -1.2 -1.1 -1.1 -1.1 1.2 -1.2 3.3 -1.0 -1.1 -1.1 -1.0<br />

2.5 h internalized -2.7 -2.0 -2.1 -2.5 -2.7 -2.2 -1.6 -1.8 -1.5 -1.3 -1.6 -2.0 -1.5 -1.4 -1.1 -1.1 1.5 -1.3 -1.0 -1.2<br />

6.5 h internalized -3.3 -2.3 -2.4 -3.4 -3.5 -2.4 -2.1 -2.3 -1.8 -1.5 -1.4 -1.8 -1.2 -1.3 1.6 1.9 -1.1 -1.2 1.1 -1.0<br />

2.5 h serum/CO 2 control -5.1 -3.0 -2.8 -3.8 -3.9 -2.9 -4.6 -4.5 -3.3 -1.8 -1.2 -1.7 -1.2 1.1 -2.4 2.3 1.2 1.2 -1.6 -1.5<br />

6.5 h serum/CO 2 control -11.2 -5.0 -3.6 -6.2 -4.8 -3.2 -7.5 -8.8 -4.4 -1.8 -1.5 -2.1 -1.3 -1.0 -2.3 1.6 2.0 1.1 -2.0 -1.6<br />

2.5 h anaerobic incubation -3.9 -2.0 -1.5 -2.1 -2.3 -1.9 -3.8 -3.5 -2.9 -1.8 -1.2 -1.5 -1.4 1.0 -2.0 2.1 1.1 1.1 -1.9 -1.9<br />

B<br />

C<br />

10.0<br />

10.0<br />

tyrS<br />

ileS<br />

leuS<br />

alaS<br />

thrS<br />

metS<br />

1.0<br />

pheS<br />

pheT<br />

serS<br />

1.0<br />

asnS<br />

gltX<br />

trpS<br />

aspS<br />

hisS<br />

0.1<br />

0.1<br />

D<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

E<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

10.0<br />

10.0<br />

1.0<br />

valS<br />

glyS<br />

cysS<br />

1.0<br />

argS<br />

proS<br />

lysS<br />

0.1<br />

0.1<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

Fig. R.5.27: tRNA synthetase gene expression.<br />

A. Overview on fold change values relative to the baseline sample “1 h serum/CO 2 control” and on significance in statistical group<br />

comparisons. The sample “6.5 h serum/CO 2 control” could not be included in statistical testing because <strong>of</strong> small group size (n = 2).<br />

B-E. Overview on mean normalized gene expression intensity. After the global normalization <strong>by</strong> inter-chip scaling and detrending,<br />

each individual gene has been normalized to the expression level <strong>of</strong> the baseline sample “1 h serum/CO 2 control” (1 h co.). Although<br />

not being continuous data, values were depicted as line graphs instead <strong>of</strong> bar charts for better facility <strong>of</strong> inspection.<br />

Twenty tRNA synthetase genes were assigned to four groups: genes repressed in at least one time point <strong>of</strong> internalization (B), genes<br />

with trend <strong>of</strong> repression in at least one time point <strong>of</strong> internalization with p* < 0.05, but an absolute fold change <strong>of</strong> less than 2 (C),<br />

genes with trend <strong>of</strong> induction in one time point <strong>of</strong> internalization (D), and genes exhibiting no change in expression (E).<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2 control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h internalization;<br />

2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h co. − 6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic incubation)<br />

After having observed the described changes in metabolic gene expression, the possible<br />

differential expression <strong>of</strong> transporter genes was addressed directly. Here, two groups could be<br />

distinguished. The first group consisted <strong>of</strong> transporter genes with repressed gene expression in at<br />

least one analyzed time point <strong>of</strong> internalized staphylococci. The urea transporter<br />

SAOUHSC_02557 belonged to this group. Noticeably, the adjacent urease operon<br />

SAOUHSC_02558 to SAOUHSC_02565 (ureABCEFGD) possessed a similar expression pattern,<br />

although not all genes were detected as differentially expressed (Fig. R.5.28).<br />

161


mean normalized intensity<br />

mean normalized intensity<br />

mean normalized intensity<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

A<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

SAOUHSC_<br />

02557 ureA ureB ureC ureE ureF ureG ureD<br />

exponential growth phase 1.6 3.0 2.2 1.9 1.5 1.4 1.2 1.2<br />

2.5 h internalized -2.6 -2.9 -3.2 -3.0 -1.9 -2.2 -1.9 -2.0<br />

6.5 h internalized -1.1 -1.7 -1.6 -1.4 -1.3 -1.6 -1.4 -1.3<br />

2.5 h serum/CO 2 control -1.8 -3.3 -3.1 -2.1 -1.2 -1.3 -1.3 -1.4<br />

6.5 h serum/CO 2 control -1.5 -2.3 -2.0 -1.1 -1.1 -1.2 -1.4 -1.2<br />

2.5 h anaerobic incubation -1.3 5.0 4.8 3.1 1.7 1.3 1.1 -1.1<br />

Fig. R.5.28: Urea transporter and urease operon gene<br />

expression.<br />

A. Overview on fold change values relative to the baseline<br />

sample “1 h serum/CO 2 control” and on significance in statistical<br />

group comparisons. The sample “6.5 h serum/CO 2 control”<br />

could not be included in statistical testing because <strong>of</strong> small<br />

group size (n = 2).<br />

B. Overview on mean normalized gene expression intensity.<br />

After the global normalization <strong>by</strong> inter-chip scaling and<br />

detrending, each individual gene has been normalized to the<br />

expression level <strong>of</strong> the baseline sample “1 h serum/CO 2 control”<br />

(1 h co.). Although not being continuous data, values were<br />

depicted as line graphs instead <strong>of</strong> bar charts for better facility <strong>of</strong><br />

inspection. (exp. − exponential growth phase; 1 h co. – 1 h<br />

serum/CO 2 control; 2.5 h int. − 2.5 h internalization; 6.5 h int. −<br />

6.5 h internalization; 2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h<br />

co. − 6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic<br />

incubation)<br />

B<br />

10.0<br />

1.0<br />

0.1<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

SAOUHSC_02557<br />

ureA<br />

ureB<br />

ureC<br />

ureE<br />

ureF<br />

ureG<br />

ureD<br />

A<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

potA potB potC potD gltS arcD opuD opuCd opuCc opuCb opuCa<br />

exponential growth phase -1.4 -1.4 -1.6 -1.4 -1.1 -1.5 1.0 -1.0 -1.1 -1.1 -1.0<br />

2.5 h internalized -7.1 -8.4 -9.2 -4.9 -2.2 -2.4 -2.1 -3.3 -3.3 -4.6 -4.2<br />

6.5 h internalized -3.9 -4.4 -4.9 -3.7 -3.0 -2.3 -1.7 -1.6 -2.1 -2.7 -3.0<br />

2.5 h serum/CO 2 control -3.3 -3.5 -4.1 -3.1 -4.4 1.4 -4.7 2.1 2.0 1.9 2.1<br />

6.5 h serum/CO 2 control -4.8 -4.3 -5.5 -2.7 -6.0 2.0 -7.0 1.2 1.0 -1.1 -1.0<br />

2.5 h anaerobic incubation -2.9 -3.5 -4.5 -3.7 -4.0 2.1 -3.9 3.2 3.7 3.7 4.6<br />

B<br />

C<br />

10.0<br />

10.0<br />

potA<br />

opuD1<br />

potB<br />

opuCd<br />

1.0<br />

potC<br />

potD<br />

1.0<br />

opuCc<br />

opuCb<br />

gltS<br />

opuCa<br />

arcD<br />

0.1<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

0.1<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

Fig. R.5.29: Spermidine/putrescine, sodium/glutamate, arginine/ornithine, glycine betaine and amino acid transporter gene<br />

expression.<br />

A. Overview on fold change values relative to the baseline sample “1 h serum/CO 2 control” and on significance in statistical group<br />

comparisons. The sample “6.5 h serum/CO 2 control” could not be included in statistical testing because <strong>of</strong> small group size (n = 2).<br />

B, C. Overview on mean normalized gene expression intensity. After the global normalization <strong>by</strong> inter-chip scaling and detrending,<br />

each individual gene has been normalized to the expression level <strong>of</strong> the baseline sample “1 h serum/CO 2 control” (1 h co.). Although<br />

not being continuous data, values were depicted as line graphs instead <strong>of</strong> bar charts for better facility <strong>of</strong> inspection.<br />

Spermidine/putrescine ABC transporter potABCD, sodium/glutamate symporter gltS, arginine/ornithine antiporter arcD (B) and<br />

glycine betaine transporter opuD1 and amino acid ABC transporter opuCdCcCbCa (C).<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2 control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h internalization;<br />

2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h co. − 6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic incubation)<br />

162


mean normalized intensity<br />

mean normalized intensity<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

Fig. R.5.30:<br />

Choline, citrate, xanthine, fructose, and lactate<br />

transporter gene expression.<br />

A. Overview on fold change values relative to the baseline<br />

sample “1 h serum/CO 2 control” and on significance in<br />

statistical group comparisons. Differentially expressed<br />

genes are marked <strong>by</strong> gray filling for the corresponding<br />

sample condition/time point.<br />

Genes which were significant in statistical testing<br />

(p* < 0.05) but did not pass the absolute fold change<br />

cut<strong>of</strong>f 2 are indicated in bold without filling. The sample<br />

“6.5 h serum/CO 2 control” could not be included in<br />

statistical testing because <strong>of</strong> small group size (n = 2).<br />

B. Overview on mean normalized gene expression<br />

intensity. After the global normalization <strong>by</strong> inter-chip<br />

scaling and detrending, each individual gene has been<br />

normalized to the expression level <strong>of</strong> the baseline sample<br />

“1 h serum/CO 2 control”. Although not being continuous<br />

data, values were depicted as line graphs instead <strong>of</strong> bar<br />

charts for better facility <strong>of</strong> inspection.<br />

Choline transporter cudT, citrate transporter<br />

SAOUHSC_02943, xanthine permease pbuX, fructose<br />

specific permease fruA, and L-lactate permease lldP2.<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2<br />

control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h<br />

internalization; 2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h<br />

co. − 6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h<br />

anaerobic incubation)<br />

A<br />

B<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

10.00<br />

1.00<br />

0.10<br />

0.01<br />

exp.<br />

cudT<br />

SAOUHSC_<br />

02943 pbuX fruA lldP2<br />

exponential growth phase 1.5 -1.2 -1.1 -4.3 -8.6<br />

2.5 h internalized -2.5 -3.7 -3.5 -3.1 -14.4<br />

6.5 h internalized -2.1 -1.4 -1.6 -2.0 -13.2<br />

2.5 h serum/CO 2 control -1.6 1.4 1.5 -9.0 1.4<br />

6.5 h serum/CO 2 control -1.6 -1.3 -1.2 -6.8 -2.5<br />

2.5 h anaerobic incubation -2.9 -1.5 -1.1 -4.7 2.4<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

cudT<br />

SAOUHSC_02943<br />

pbuX<br />

fruA<br />

lldP2<br />

Furthermore, expression <strong>of</strong> spermidine/putrescine ABC transporter potABCD,<br />

sodium/glutamate symporter gltS, arginine/ornithine antiporter arcD, glycine betaine transporter<br />

opuD1, and amino acid ABC transporter opuCdCcCbCa genes was repressed in internalized<br />

staphylococci (Fig. R.5.29). Additional examples <strong>of</strong> repressed transporter genes were the choline<br />

transporter cudT, the citrate transporter SAOUHSC_02943, the xanthine permease pbuX, the<br />

fructose specific permease fruA, and the L-lactate permease lldP2 (Fig. R.5.30).<br />

A<br />

Fig. R.5.31: Phosphate ABC transporter gene<br />

expression.<br />

A. Overview on fold change values relative to the<br />

baseline sample “1 h serum/CO 2 control” and on<br />

significance in statistical group comparisons.<br />

Differentially expressed genes are marked <strong>by</strong> gray<br />

filling for the corresponding sample condition/time<br />

point.<br />

Genes which were significant in statistical testing<br />

(p* < 0.05) but did not pass the absolute fold change<br />

cut<strong>of</strong>f 2 are indicated in bold without filling. The<br />

sample “6.5 h serum/CO 2 control” could not be<br />

included in statistical testing because <strong>of</strong> small group<br />

size (n = 2).<br />

B. Overview on mean normalized gene expression<br />

intensity. After the global normalization <strong>by</strong> inter-chip<br />

scaling and detrending, each individual gene has been<br />

normalized to the expression level <strong>of</strong> the baseline<br />

sample “1 h serum/CO 2 control”. Although not being<br />

continuous data, values were depicted as line graphs<br />

instead <strong>of</strong> bar charts for better facility <strong>of</strong> inspection.<br />

(exp. − exponential growth phase; 1 h co. – 1 h<br />

serum/CO 2 control; 2.5 h int. − 2.5 h internalization;<br />

6.5 h int. − 6.5 h internalization; 2.5 h co. − 2.5 h<br />

serum/CO 2 control; 6.5 h co. − 6.5 h serum/CO 2 control;<br />

2.5 h anae. − 2.5 h anaerobic incubation)<br />

B<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

100.0<br />

10.0<br />

1.0<br />

0.1<br />

exp.<br />

phoU pstB pstA pstC SAOUHSC_<br />

01388<br />

pstS<br />

exponential growth phase 1.3 1.1 -1.5 -1.5 -1.9 -3.2<br />

2.5 h internalized 5.6 10.1 16.8 43.7 47.7 38.8<br />

6.5 h internalized 9.8 16.1 27.4 59.6 61.3 34.0<br />

2.5 h serum/CO 2 control -1.3 -1.4 -1.5 -1.1 -1.0 -2.5<br />

6.5 h serum/CO 2 control -1.7 -1.4 -1.7 -1.3 -1.7 -2.2<br />

2.5 h anaerobic incubation -1.2 -1.6 -1.9 -1.5 -1.8 1.2<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

phoU<br />

pstB<br />

pstA<br />

pstC<br />

SAOUHSC _01388<br />

pstS<br />

163


mean normalized intensity<br />

mean normalized intensity<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

The second group comprised transporter genes with induced gene expression in at least one<br />

analyzed time point <strong>of</strong> internalized staphylococci. This applied for example to the phosphate ABC<br />

transporter operon SAOUHSC_01384 to SAOUHSC_01389, pho/pst (Fig. R.5.31).<br />

A<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

phnE<br />

SAOUHSC_<br />

00103 phnC SAOUHSC_<br />

00105 metN1 SAOUHSC_<br />

00424<br />

SAOUHSC_<br />

00426 metN2<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

exponential growth phase -1.4 -1.5 -1.6 -1.2 3.3 2.8 2.0 1.6<br />

2.5 h internalized -1.1 1.2 2.0 4.8 6.6 5.8 3.6 1.6<br />

6.5 h internalized 2.9 3.6 5.2 7.4 16.6 12.8 7.6 2.4<br />

2.5 h serum/CO 2 control -1.0 1.1 1.3 1.5 48.7 39.8 22.7 3.1<br />

6.5 h serum/CO 2 control -1.4 -1.4 -1.1 -1.4 21.6 20.1 10.5 2.4<br />

2.5 h anaerobic incubation -1.5 -1.1 1.1 1.7 3.9 2.8 1.8 1.4<br />

Fig. R.5.32: Phosphonate transporter and methionine<br />

binding protein gene expression.<br />

A. Overview on fold change values relative to the baseline<br />

sample “1 h serum/CO 2 control” and on significance in<br />

statistical group comparisons. The sample “6.5 h serum/CO 2<br />

control” could not be included in statistical testing because<br />

<strong>of</strong> small group size (n = 2).<br />

B. Overview on mean normalized gene expression intensity.<br />

After the global normalization <strong>by</strong> inter-chip scaling and<br />

detrending, each individual gene has been normalized to<br />

the expression level <strong>of</strong> the baseline sample “1 h serum/CO 2<br />

control” (1 h co.). Although not being continuous data,<br />

values were depicted as line graphs instead <strong>of</strong> bar charts for<br />

better facility <strong>of</strong> inspection.<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2<br />

control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h<br />

internalization; 2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h<br />

co. − 6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic<br />

incubation)<br />

B<br />

100.0<br />

10.0<br />

1.0<br />

0.1<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

phnE<br />

SAOUHSC_00103<br />

phnC<br />

SAOUHSC_00105<br />

metN1<br />

SAOUHSC_00424<br />

SAOUHSC_00426<br />

metN2<br />

Induction was detected for phosphonate ABC transporters (operon SAOUHSC_00102 to<br />

SAOUHSC_00104, including phnE and phnC, and the adjacent SAOUHSC_00105), methionine<br />

import ATP-binding protein gene metN1 and transporter/permease SAOUHSC_00424 and<br />

SAOUHSC_00426 from the same operon, and the separately encoded methionine import ATPbinding<br />

protein gene metN2 (Fig. R.5.32).<br />

A<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

oppF<br />

SAOUHSC_<br />

00169<br />

SAOUHSC_<br />

00923<br />

oppC oppD SAOUHSC_<br />

00926<br />

oppA<br />

SAOUHSC_<br />

00928<br />

appD appF SAOUHSC_<br />

00931<br />

appC<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

exponential growth phase 1.3 1.2 1.4 1.4 1.4 1.3 1.3 1.3 1.3 -1.1 1.5 1.5<br />

2.5 h internalized 1.3 5.2 2.7 2.7 3.0 2.4 2.4 1.2 -1.0 -1.5 -1.2 -1.2<br />

6.5 h internalized 2.2 17.8 4.8 4.9 5.0 4.3 4.4 2.8 1.9 1.2 1.5 1.0<br />

2.5 h serum/CO 2 control 4.9 27.6 8.3 8.2 9.0 7.5 7.6 8.2 3.1 1.7 1.6 1.3<br />

6.5 h serum/CO 2 control 3.4 20.6 4.2 5.2 5.7 5.5 5.6 10.2 4.1 2.4 1.8 1.3<br />

2.5 h anaerobic incubation 1.2 1.8 5.4 5.1 6.0 4.6 4.8 3.4 1.5 -1.0 -1.0 -1.0<br />

Fig. R.5.33: Peptide and oligopeptide transporter gene<br />

expression.<br />

A. Overview on fold change values relative to the baseline<br />

sample “1 h serum/CO 2 control” and on significance in<br />

statistical group comparisons. The sample “6.5 h serum/CO 2<br />

control” could not be included in statistical testing because<br />

<strong>of</strong> small group size (n = 2).<br />

B. Overview on mean normalized gene expression intensity.<br />

After the global normalization <strong>by</strong> inter-chip scaling and<br />

detrending, each individual gene has been normalized to<br />

the expression level <strong>of</strong> the baseline sample “1 h serum/CO 2<br />

control” (1 h co.). Although not being continuous data,<br />

values were depicted as line graphs instead <strong>of</strong> bar charts for<br />

better facility <strong>of</strong> inspection.<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2<br />

control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h<br />

internalization; 2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h<br />

co. − 6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic<br />

incubation)<br />

B<br />

100.0 oppF<br />

10.0<br />

1.0<br />

0.1<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

SAOUHSC_00169<br />

SAOUHSC_00923<br />

oppC<br />

oppD<br />

SAOUHSC_00926<br />

oppA<br />

SAOUHSC_00928<br />

appD<br />

appF<br />

SAOUHSC_00931<br />

appC<br />

164


mean normalized intensity<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

Additionally, a high number <strong>of</strong> peptide/oligopeptide transporters were induced: peptide<br />

transporters oppF and SAOUHSC_00169, oligopeptide transporters SAOUHSC_00926 and oppA,<br />

and genes SAOUHSC_00923, oppC (SAOUHSC_00924), and oppD (SAOUHSC_00925) from the<br />

same operon (SAOUHSC_00923 to SAOUHSC_00927). Also induced SAOUHSC_00928 gene codes<br />

for an oligopeptide ABC transporter. In the same chromosomal region, a four-gene-operon<br />

(SAOUHSC_00929 to SAOUHSC_00932) encodes oligopeptide ABC transporters appD, appF,<br />

SAOUHSC_00931, and appC, but these were not induced in internalized staphylococci<br />

(Fig. R.5.33).<br />

Further example for induced transporter gene expression was another operon including msmX<br />

(SAOUHSC_00175, multiple sugar-binding transport ATP-binding protein), SAOUHSC_00176<br />

(bacterial extracellular solute-binding protein), SAOUHSC_00177 and SAOUHSC_00178 (both<br />

maltose ABC transporter, permease proteins). Also opuD2 (osmoprotectant transporter), glnQ<br />

(amino acid ABC transporter), SAOUHSC_02729 (amino acid ABC transporter-like protein), and<br />

gntP (gluconate permease) were induced (Fig. R.5.34).<br />

A<br />

fold change in comparison to<br />

baseline 1 h serum/CO 2 control<br />

S. aureus<br />

RN1HG<br />

sample<br />

conditions<br />

and time<br />

points<br />

msmX SAOUHSC_<br />

00176<br />

SAOUHSC_<br />

00177<br />

SAOUHSC_<br />

00178<br />

opuD2 glnQ<br />

SAOUHSC_<br />

02729<br />

exponential growth phase -1.3 -1.3 -1.2 -1.0 2.8 1.5 1.0 -1.0<br />

2.5 h internalized 15.6 7.6 3.5 2.7 3.7 2.1 8.4 2.6<br />

6.5 h internalized 8.2 5.6 2.8 1.8 2.9 -1.0 2.9 2.1<br />

2.5 h serum/CO 2 control 1.4 -1.1 1.1 1.0 2.9 -1.2 -1.2 -1.2<br />

6.5 h serum/CO 2 control 43.7 42.4 34.8 29.5 4.2 1.1 7.2 7.0<br />

2.5 h anaerobic incubation 3.5 2.3 1.7 1.7 1.1 -1.1 1.2 -1.0<br />

gntP<br />

Fig. R.5.34: Sugar/maltose, osmoprotectant, amino acid,<br />

and gluconate transporter or permease gene expression.<br />

A. Overview on fold change values relative to the baseline<br />

sample “1 h serum/CO 2 control” and on significance in<br />

statistical group comparisons. The sample “6.5 h serum/CO 2<br />

control” could not be included in statistical testing because<br />

<strong>of</strong> small group size (n = 2).<br />

B. Overview on mean normalized gene expression intensity.<br />

After the global normalization <strong>by</strong> inter-chip scaling and<br />

detrending, each individual gene has been normalized to<br />

the expression level <strong>of</strong> the baseline sample “1 h serum/CO 2<br />

control” (1 h co.). Although not being continuous data,<br />

values were depicted as line graphs instead <strong>of</strong> bar charts for<br />

better facility <strong>of</strong> inspection.<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2<br />

control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h<br />

internalization; 2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h<br />

co. − 6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic<br />

incubation)<br />

B<br />

100.0<br />

10.0<br />

1.0<br />

0.1<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

msmX<br />

SAOUHSC_00176<br />

SAOUHSC_00177<br />

SAOUHSC_00178<br />

opuD2<br />

glnQ<br />

SAOUHSC_02729<br />

gntP<br />

Gene expression changes in internalized staphylococci not occurring in the same way in any <strong>of</strong><br />

the control samples<br />

Analysis <strong>of</strong> virulence associated or metabolic gene expression changes revealed a common<br />

aspect <strong>of</strong> the internalization signature: For many genes, some <strong>of</strong> the control conditions or time<br />

points exhibited a similar expression change as the internalized expression pattern. To get hints<br />

on real specifically internalization dependent gene expression changes, a set <strong>of</strong> sequences, which<br />

contained sequences with different expression in internalized staphylococci in comparison to all<br />

other control samples, was defined in two steps. First, subsets <strong>of</strong> regulated sequences in<br />

internalized staphylococci were constituted which were a) not regulated in 2.5 h serum/CO 2<br />

control and b) not regulated in 2.5 h anaerobically incubated samples (each in reference to the<br />

baseline sample <strong>of</strong> 1 h serum/CO 2 control with p* < 0.05 and a minimal absolute fold change <strong>of</strong><br />

2). In these sets, sequences were still included which possessed differing expression in the 6.5 h<br />

serum/CO 2 control in comparison to the baseline samples.<br />

165


mean normalized intensity<br />

mean normalized intensity<br />

mean normalized intensity<br />

mean normalized intensity<br />

mean normalized intensity<br />

mean normalized intensity<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

A<br />

B<br />

45 37 34<br />

98 75 43<br />

C<br />

internalization-specific<br />

differential expression;<br />

induction 2.5 h after<br />

start <strong>of</strong> infection<br />

internalization-specific<br />

differential expression;<br />

induction 6.5 h after<br />

start <strong>of</strong> infection<br />

F<br />

internalization-specific<br />

differential expression;<br />

repression 2.5 h after<br />

start <strong>of</strong> infection<br />

internalization-specific<br />

differential expression;<br />

repression 6.5 h after<br />

start <strong>of</strong> infection<br />

10E+02<br />

10E+02<br />

10E+01 10.00<br />

10E+01 10.00<br />

10E+00 1.00<br />

10E+00 1.00<br />

10E-01 0.10<br />

10E-01 0.10<br />

D<br />

10E-02 0.01<br />

1 2 3 4 5 6 7<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

G<br />

10E-02 0.01<br />

1 2 3 4 5 6 7<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

10E+02<br />

10E+02<br />

10E+01 10.00<br />

10E+01 10.00<br />

10E+00 1.00<br />

10E+00 1.00<br />

10E-01 0.10<br />

10E-01 0.10<br />

E<br />

10E-02 0.01<br />

10E-02 0.01<br />

10E+02<br />

1 2 3 4 5 6 7<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

H<br />

10E+02<br />

1 2 3 4 5 6 7<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

10E+01 10.00<br />

10E+01 10.00<br />

10E+00 1.00<br />

10E+00 1.00<br />

10E-01 0.10<br />

10E-01 0.10<br />

10E-02 0.01<br />

10E-02 0.01<br />

1 2 3 4 5 6 7<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

1 2 3 4 5 6 7<br />

exp.<br />

1 h<br />

co.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

2.5 h<br />

anae.<br />

S. aureus RN1HG sample conditions and time points<br />

Fig. R.5.35: Internalization-specific differential gene expression.<br />

A, B. Overview on the number <strong>of</strong> sequences which were differentially expressed in internalized staphylococci in comparison to all<br />

other control samples and exhibited induction (A) or repression (B).<br />

C-H. Overview on mean normalized gene expression intensity. After the global normalization <strong>by</strong> inter-chip scaling and detrending,<br />

each individual gene has been normalized to the expression level <strong>of</strong> the baseline sample “1 h serum/CO 2 control” (1 h co.). Although<br />

not being continuous data, values were depicted as line graphs instead <strong>of</strong> bar charts for better facility <strong>of</strong> inspection.<br />

Intensity values are depicted for 45 sequences specifically induced after 2.5 h (C), 37 sequences specifically induced at both time<br />

points (D), 34 sequences specifically induced after 6.5 h (E), 98 sequences specifically repressed after 2.5 h (F), 75 sequences<br />

specifically repressed at both time points (G), and 43 sequences specifically repressed after 6.5 h (H).<br />

(exp. − exponential growth phase; 1 h co. – 1 h serum/CO 2 control; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h internalization;<br />

2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h co. − 6.5 h serum/CO 2 control; 2.5 h anae. − 2.5 h anaerobic incubation)<br />

166


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

This difference could not be assessed with statistical testing because only two arrays were<br />

available for the 6.5 h control. Thus, in a second step, all sequences which possessed an absolute<br />

fold change equal or greater than 2 in the comparison between 6.5 h and 1 h serum/CO 2 control<br />

were excluded from the subsets generated in the first step. This procedure was performed for<br />

induced and repressed genes separately and resulted in 45 sequences specifically induced after<br />

2.5 h, 37 sequences specifically induced at both time points, 34 sequences specifically induced<br />

after 6.5 h, 98 sequences specifically repressed after 2.5 h, 75 sequences specifically repressed at<br />

both time points, and 43 sequences specifically repressed after 6.5 h in samples <strong>of</strong> internalized<br />

staphylococci (Fig. R.5.35). Most interestingly, these lists included besides newly identified<br />

transcripts also genes which were described above in the context <strong>of</strong> virulence or metabolic<br />

genes, e. g. clfB, coa, phoP, prsA, glpK (2.5 h, induced), fnbB, vwb, ssaA, phoU, pstA, pstB, pstC,<br />

pstS (2.5 h and 6.5 h, induced), chp, lukD, lukE, hla, efb, emp (6.5 h, induced), icaD, icaC, rot, nuc,<br />

sspA, urea transporter SAOUHSC_02557, ureF (2.5 h, repressed), sarU (6.5 h, repressed).<br />

New transcripts identified with the tiling array approach which are regulated in S9 cell<br />

internalized staphylococci<br />

All sequences <strong>of</strong> the tiling arrays were included in the statistical testing to identify the<br />

internalization specific gene expression signature. Therefore, the resulting lists <strong>of</strong> differentially<br />

expressed genes contained newly identified transcripts: 200 sequences for the 2.5 h and 138<br />

sequences for the 6.5 h time point which were significant in ANOVA with p* < 0.05 and exhibited<br />

a minimal absolute fold change <strong>of</strong> 2 (Table R.5.5).<br />

When comparing these lists, 100 transcripts were differentially expressed at both time points<br />

(Fig. R.5.36 A). These consisted <strong>of</strong> 67 induced and 33 repressed sequences (Fig. R.5.36 B).<br />

A<br />

B<br />

100 100 38<br />

2.5 h<br />

repressed<br />

repressed 6.5 h<br />

induced<br />

35 18<br />

induced<br />

new transcripts differentially<br />

expressed in the comparison<br />

“2.5 h internalization” vs.<br />

“1 h serum/CO 2 control”<br />

new transcripts differentially<br />

expressed in the comparison<br />

“6.5 h internalization” vs.<br />

“1 h serum/CO 2 control”<br />

65<br />

33<br />

67<br />

20<br />

Fig. R.5.36: Comparison <strong>of</strong> newly identified transcripts in the 2.5 h and 6.5 h signatures <strong>of</strong> internalized staphylococci.<br />

Both internalized samples were compared to the baseline <strong>of</strong> 1 h serum/CO 2 control with statistical testing and multiple testing<br />

correction (p* < 0.05), and a minimal absolute fold change cut<strong>of</strong>f <strong>of</strong> 2 was applied. The comparison <strong>of</strong> differentially expressed newly<br />

identified transcripts at the 2.5 h and 6.5 h time point was performed for all regulated transcripts (A) and for induced and repressed<br />

transcripts separately (B).<br />

Although these 200 and 138 newly identified transcripts were known to be differentially<br />

expressed in at least one <strong>of</strong> the two internalized samples in comparison to the 1 h serum/CO 2<br />

control, they were expected to form further subgroups e. g. dependent on the direction <strong>of</strong><br />

regulation. Therefore, a k-means clustering was applied to a ratio data set (experimental<br />

condition normalized to the 1 h serum/CO 2 control) consisting <strong>of</strong> the five experimental conditions<br />

<strong>of</strong> non-adherent (1 h), internalized (2.5 h and 6.5 h), and serum/CO 2 control (2.5 h and 6.5 h)<br />

staphylococci.<br />

167


log 2 (ratio)<br />

log 2 (ratio)<br />

log 2 (ratio)<br />

log 2 (ratio)<br />

log 2 (ratio)<br />

log 2 (ratio)<br />

log 2 (ratio)<br />

log 2 (ratio)<br />

log 2 (ratio)<br />

log 2 (ratio)<br />

log 2 (ratio)<br />

log 2 (ratio)<br />

Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

For the new transcripts included in the 2.5 h internalization signature, 6 cluster groups were<br />

defined (Fig. R.5.37 A). Cluster group 1 contained 26 transcripts. Expression was increased in the<br />

2.5 h samples and still slightly stronger in the 6.5 h samples. The serum/CO 2 controls at both time<br />

points also exhibited increase <strong>of</strong> expression. Cluster group 2 harbored 32 transcripts which<br />

featured increase in expression at the 2.5 h and to a lesser extent at the 6.5 h internalization time<br />

points and in the later 6.5 h serum/CO 2 control whereas the 2.5 h serum/CO 2 control exhibited in<br />

average no change. Transcripts with decreased expression in the 2.5 h internalized and less<br />

strongly in the 6.5 h sample belonged to cluster group 3 which included 42 transcripts. The 2.5 h<br />

serum/CO 2 control exhibited no apparent trend <strong>of</strong> regulation whereas the 6.5 h control exhibited<br />

a decrease in expression. In cluster group 4, 26 transcripts were classified, which exhibited a<br />

decrease <strong>of</strong> expression in both internalization samples, a slight increase in the 2.5 h serum/CO 2<br />

controls and a decline <strong>of</strong> expression to the reference level in the 6.5 h serum/CO 2 controls.<br />

Internalization samples in cluster group 5 showed an increase <strong>of</strong> expression 2.5 h after start <strong>of</strong><br />

infection, which again declined in samples after 6.5 h. Contrarily, expression in the serum/CO 2<br />

controls was induced at the 2.5 h time point and further increased at the 6.5 h time point. This<br />

cluster contained 49 sequences. Finally, cluster group 6 contained 25 sequences which were<br />

increased in both internalized samples, and possessed a variable expression pattern in the 2.5 h<br />

serum/CO 2 control with increased, repressed and unchanged values, while the 6.5 h serum/CO 2<br />

control samples exhibited a trend <strong>of</strong> decreased expression.<br />

A<br />

B<br />

cluster group 1<br />

cluster group 2<br />

cluster group 1<br />

cluster group 2<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

1 h<br />

n. ad.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

1 h<br />

n. ad.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

1 h<br />

n. ad.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

1 h<br />

n. ad.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

cluster group 3<br />

cluster group 4<br />

cluster group 3<br />

cluster group 4<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

1 h<br />

n. ad.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

1 h<br />

n. ad.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

1 h<br />

n. ad.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

1 h<br />

n. ad.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

cluster group 5<br />

cluster group 6<br />

cluster group 5<br />

cluster group 6<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

1 h<br />

n. ad.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

1 h<br />

n. ad.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

1 h<br />

n. ad.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

6<br />

4<br />

2<br />

0<br />

-2<br />

-4<br />

-6<br />

1 h<br />

n. ad.<br />

2.5 h<br />

int.<br />

6.5 h<br />

int.<br />

2.5 h<br />

co.<br />

6.5 h<br />

co.<br />

Fig. R.5.37: K-means clustering <strong>of</strong> gene expression data from newly identified transcripts, which exhibited differential expression in<br />

the 2.5 h internalized sample (A) or in the 6.5 h internalized sample (B) in comparison to the 1 h serum/CO 2 control with statistical<br />

testing and multiple testing correction (p* < 0.05) and a minimal absolute fold change cut<strong>of</strong>f <strong>of</strong> 2. For k-means clustering, the<br />

following parameters were applied: Transcripts should be assigned to the user-defined number <strong>of</strong> 6 groups and sorted <strong>by</strong> deviation<br />

using the metric type cosine correlation. No additional data preparation or statistical cuts were applied.<br />

The ratio data set (experimental condition normalized to the 1 h serum/CO 2 control) employed for clustering consisted <strong>of</strong> the five<br />

experimental conditions <strong>of</strong> non-adherent (1 h), internalized (2.5 h and 6.5 h), and serum/CO 2 control (2.5 h and 6.5 h) staphylococci.<br />

Although not being continuous data, values were depicted as line graphs instead <strong>of</strong> bar charts for better facility <strong>of</strong> inspection.<br />

(1 h n. ad. – non-adherent staphylococci after 1 h <strong>of</strong> infection; 2.5 h int. − 2.5 h internalization; 6.5 h int. − 6.5 h internalization;<br />

2.5 h co. − 2.5 h serum/CO 2 control; 6.5 h co. − 6.5 h serum/CO 2 control)<br />

Correspondingly, also for the new transcripts included in the 6.5 h internalization signature,<br />

6 cluster groups were defined (Fig. R.5.37 B). Here, cluster group 1 contained 19 transcripts with<br />

168


Maren Depke<br />

Results<br />

Pathogen Gene Expression Pr<strong>of</strong>iling<br />

an increase in expression at both internalization time points and in the later 6.5 h serum/CO 2<br />

control whereas the 2.5 h serum/CO 2 control exhibited in average no change. In cluster group 2,<br />

to which 17 sequences were assigned, transcripts exhibited a decrease <strong>of</strong> expression in both<br />

internalization samples, and in general no change in the 2.5 h serum/CO 2 controls. Some <strong>of</strong> the<br />

sequences were slightly decreased and some other slightly increased in the later 6.5 h serum/CO 2<br />

controls. Cluster group 3 included 36 transcripts with increased expression at both internalization<br />

and serum/CO 2 control time points. Transcripts with mainly decreased expression in both<br />

internalized samples belonged to cluster group 4 which included 34 transcripts. The 2.5 h<br />

serum/CO 2 control exhibited a varying trend <strong>of</strong> regulation with some repressed and some<br />

induced sequences, whereas the 6.5 h control exhibited a decrease in expression. The 16<br />

transcripts in cluster group 5 possessed an increase in the 2.5 h and 6.5 h internalization samples.<br />

Also in the serum/CO 2 controls at both time points a slight tendency <strong>of</strong> induction <strong>of</strong> expression<br />

was observed, but less prominent than the internalization samples. The last cluster group 6<br />

comprised the remaining 16 transcripts with similar expression pattern as the transcripts in<br />

cluster group 5 with exception <strong>of</strong> the expression change in the serum/CO 2 controls, where only<br />

minor changes in the 2.5 h and slight decrease in expression in the 6.5 h samples were visible.<br />

Transcripts in all cluster groups <strong>of</strong> both 2.5 h and 6.5 h internalization signatures showed only<br />

minor variation from the non-changed value <strong>of</strong> 1 in the sample <strong>of</strong> non-adherent staphylococci.<br />

This finding was expected as the similarity between this sample and the 1 h serum/CO 2 control<br />

baseline had been unveiled before.<br />

169


Maren Depke<br />

D I S C U S S I O N A N D C O N C L U S I O N S<br />

LIVER GENE EXPRESSION PATTERN IN A MOUSE<br />

PSYCHOLOGICAL STRESS MODEL<br />

BALB/c mice were subjected to combined acoustic and restraint stress for 2 h twice a day<br />

during a period <strong>of</strong> 4.5 days, which serves as a murine model <strong>of</strong> chronic psychological stress.<br />

Experiments using the same model had already revealed that mice suffered from an impaired<br />

antibacterial defense and from depression-like behavior (Kiank et al. 2006, 2007a).<br />

The observation <strong>of</strong> severe loss <strong>of</strong> total body mass initiated further analysis <strong>of</strong> metabolic<br />

processes as well as a hepatic gene expression pr<strong>of</strong>iling study, which resulted in evidence for the<br />

development <strong>of</strong> a hypermetabolic syndrome in chronically stressed mice. The results <strong>of</strong> liver gene<br />

expression pr<strong>of</strong>iling and physiological analyses have been published <strong>by</strong> Depke et al. in 2008 and<br />

2009.<br />

Already a single acute stress exposure caused pr<strong>of</strong>ound changes in hepatic gene expression.<br />

Genes important for metabolic pathways regulating the carbohydrate turnover showed stressinduced<br />

alterations <strong>of</strong> mRNA expression in hepatic tissue. An acute stress response is essential<br />

for energy mobilization to “fight or flight” in a potentially harmful situation and to sustain or<br />

reconstitute allostasis (McEwen 2004). Initially, catecholamines activate glycogenolysis,<br />

gluconeogenesis, and accelerate lipolysis that subsequently is assisted <strong>by</strong> catabolic<br />

glucocorticoid-induced pathways (Bag<strong>by</strong> et al. 1992, Leibowitz/Wortley 2004, Lundberg 2005,<br />

McGuinness et al. 1999). Acute psychological stress was associated with activation <strong>of</strong> the stress<br />

axes, and an induction <strong>of</strong> the expression <strong>of</strong> gluconeogenic genes and <strong>of</strong> transporters for<br />

glucogenic amino acids (Pck1, G6pc, Slc37a4, Slc15a4, Slc25a15, Slc38a2, Slc3a1, Sds, Slc6a6, and<br />

Tat).<br />

Despite the observed gene expression changes the metabolic parameters did not change<br />

significantly when stress was restricted to a singular event. Contrarily, when stress was<br />

repeatedly applied, female BALB/c mice developed severe systemic neuroendocrine and<br />

metabolic alterations.<br />

Several researchers found that repeated restraint stress causes a loss <strong>of</strong> body weight in<br />

rodents, which was mainly mediated <strong>by</strong> initially increased energy expenditure and reduced food<br />

intake that normalized or even heightened within a few days after starting repeated stress due to<br />

neuroendocrine adaptations (Harris et al. 2002, 2006). Others showed prolonged lowered food<br />

intake during 4.5 days repeated restraint stress due to prolonged neuroendocrine dysregulation<br />

(Ricart-Jané et al. 2002). In the chronic stress model used in this study no changes <strong>of</strong> total food<br />

171


Maren Depke<br />

Discussion and Conclusions<br />

consumption during 4.5 days in repeatedly stressed animals compared with nonstressed controls<br />

were observed. Unchanged consumption may result from an initially reduced food intake after<br />

the first stress session and increased food intake in the later phase <strong>of</strong> repeated stress exposure in<br />

which additionally neuroendocrine and metabolic dysregulation were manifested.<br />

Repeatedly stressed mice suffered from an increase in metabolic rate in excess <strong>of</strong> the normal<br />

metabolic response. Such a hypermetabolic response leads to a marked increase in energy<br />

demands. Protein inappropriately becomes an energy source, and increased use <strong>of</strong> protein<br />

rapidly depletes lean body mass. A hypercatabolic response is a typical feature in infection,<br />

cancer, and prolonged critical illness, and goes along with fever, dysregulations <strong>of</strong> the<br />

cardiovascular system, hyperglycemia, dyslipidemia, accelerated proteolysis, tissue damage/cell<br />

death, perfusion disturbances, and invasion <strong>by</strong> microorganisms (Alberda et al. 2006, Costelli et al.<br />

1993, Mizock 1995, Morley et al. 2006, Vanhorebeek/Van den Berghe 2004, van Waardenburg et<br />

al. 2006, Wilmore 2000, Wray et al. 2006). In contrast, starvation is connected with diminished<br />

food intake, hyperthyroidism, and reduced protein catabolism (Alberda et al. 2006, Morley et al.<br />

2006). During a hypercatabolic response, as it was detected in the model <strong>of</strong> repeated stress, food<br />

intake is <strong>of</strong>ten normal (Alberda et al. 2006, Morley et al. 2006) and associated with<br />

hypothyroidism (Vanhorebeek/Van den Berghe 2004). Recently, it was shown that prolonged<br />

sleep deprivation also can cause such a hypermetabolic response in rats (Everson/Reed 1995,<br />

Koban/Swinson 2005). In this study, the stress experiments were performed in the recovery<br />

phase <strong>of</strong> the animals, and sleep deprivation may have affected metabolic functions. Both<br />

repeatedly stressed mice and long-term sleep deprived rats showed lowered total T 3 and T 4 levels<br />

that did not depend on altered TSH concentrations (Koban/Swinson 2005). Koban and Swinson<br />

propose a reciprocal relationship <strong>of</strong> catecholamines, which progressively increase during sleep<br />

deprivation, and thyroid hormone concentrations that decline continuously in prolonged<br />

reduction <strong>of</strong> sleep time. However, in contrast to the repeatedly stressed mice that showed<br />

unaltered food intake in this study, sleep-deprived rats were hyperphagic while body mass was<br />

massively consumed (Everson/Reed 1995, Koban/Swinson 2005). Sleep deprived rats did not<br />

exhibit altered glucocorticoid levels, whereas chronic psychological stress characteristically was<br />

associated with persistently high corticosterone concentrations in the plasma.<br />

The activation <strong>of</strong> the central nervous system can pr<strong>of</strong>oundly affect metabolic regulation such<br />

as shown for the thyroid hormone release (Koban/Swinson 2005). The target organ <strong>of</strong> metabolic<br />

regulatory pathways is predominantly the liver (Dallman et al. 2007, Leibowitz/Wortley 2004,<br />

McEwen 2004). The activation <strong>of</strong> the HPA axis with increased glucocorticoid levels can stimulate<br />

food intake and activate carbohydrate, fat and protein catabolism. In turn, the brain receives<br />

signals such as actual glucose and lipid concentrations or increased energy demand (Lam et al.<br />

2007, Lundberg 2005, McEwen 2004). In consequence, central nervous system activation induces<br />

regulatory pathways that equilibrate metabolism to supply the needed energy, e. g. increasing<br />

glucose formation in the liver (Lam et al. 2007, Leibowitz/Wortley 2004, McEwen 2004).<br />

Hypercortisolism shifted metabolic functions toward carbohydrate, lipid, and protein<br />

catabolism (Aikawa et al. 1972, Dallman et al. 2007, Harris et al. 2002, Lam et al. 2007, Ricart-<br />

Jané et al. 2002, Souba et al. 1985) to sustain energy supply <strong>by</strong> replenishing glucose that is the<br />

main energy source <strong>of</strong> the body. Glucose can be released after glycogenolysis or induction <strong>of</strong><br />

gluconeogenesis when the supply with food is insufficient. Primarily, alanine and glutamate are<br />

precursor molecules for gluconeogenesis (Aikawa et al. 1972, Brosnan 2000, Hagopian et al.<br />

2003, Pasini et al. 2004, Souba et al. 1985). Hepatic induction <strong>of</strong> alanine aminotransferase (Gpt) 2<br />

172


Maren Depke<br />

Discussion and Conclusions<br />

and Got1 may supply the metabolites for glucose synthesis (Aikawa et al. 1972, Pasini et al. 2004,<br />

Souba et al. 1985). Thereafter, alanine and glutamate can be reconstituted <strong>by</strong> biotransformation,<br />

whereas essential amino acids cannot be replenished <strong>by</strong> biosynthesis in repeatedly stressed mice<br />

(Brosnan 2000, Hagopian et al. 2003). Deamination <strong>of</strong> amino acids results in the production <strong>of</strong><br />

ammonia, which is detoxified in the liver <strong>by</strong> the urea cycle. Increased expression <strong>of</strong> the urea cycle<br />

enzyme Asl in the liver <strong>of</strong> repeatedly stressed mice along with usage <strong>of</strong> arginine and citrulline as<br />

intermediate products <strong>of</strong> the urea cycle indicates heightened stress-induced ammonia<br />

detoxification to provide C-bodies <strong>of</strong> amino acids for metabolic pathways such as<br />

gluconeogenesis. Lactate, which alternatively can serve as substrate for hepatic gluconeogenesis,<br />

was produced in high amounts in peripheral tissues during hypermetabolism such as during<br />

sepsis and can cause lactic acidosis (Aikawa et al. 1972, Banerjee et al. 2004, Capes et al. 2000,<br />

Mizock 1995). Acidosis which was detectable in repeatedly stressed mice is <strong>of</strong>ten paralleled with<br />

insulin resistance and hyperglycemia (Capes et al. 2000, Vanhorebeek/Van den Berghe 2004,<br />

van Waardenburg et al. 2006). In fact, in repeatedly stressed mice concentrations <strong>of</strong> the<br />

adipokine resistin increased. Resistin is is inducible <strong>by</strong> glucocorticoids, prolactin, and growth<br />

hormone, and has been identified to lower insulin sensitivity (Hagopian et al. 2003). Prolonged<br />

hyperglycemia, especially in critically ill patients, increases the risk <strong>of</strong> infectious complications,<br />

neuronal damage, and multiorgan dysfunction syndrome (Banerjee et al. 2004, Harris et al. 2006,<br />

van Waardenburg et al. 2006, Wray et al. 2002). In line with this, repeatedly stressed mice<br />

suffered from a reduced antimicrobial response (Kiank et al. 2006, 2007a). Lam et al. showed in<br />

2007 that increased glucose sensing <strong>by</strong> the brain and elevated intracerebral concentrations <strong>of</strong><br />

lactate reduced the hepatic secretion <strong>of</strong> VLDL cholesterol and caused a decline <strong>of</strong> plasma<br />

triglyceride levels in rodents. In addition, Ricart-Jané et al. (2002) found that repeated restraint<br />

stress caused a loss <strong>of</strong> plasmatic triacylglycerols with VLDL levels, which was accompanied <strong>by</strong> a<br />

reduced food intake. Heightened triglyceride clearance and increased lipolysis to assemble C3<br />

bodies for gluconeogenesis can result in essential fatty acid deficiency (Akhtar et al. 2005,<br />

Nemoto/Sakurai 1995, Rizki et al. 2006).<br />

Hypotriglyceridemia was detected in repeatedly stressed mice, which went along with<br />

hypercholesteremia and up-regulation <strong>of</strong> glucocorticoid-sensitive cytochrome genes in the liver<br />

(Akhtar et al. 2005, Li-Hawkins et al. 2000, Nemoto/Sakurai 1995). Cyp4a enzymes are involved in<br />

the removal <strong>of</strong> fatty acids and can counter-regulate hepatic steatosis, which was a typical<br />

consequence <strong>of</strong> repeated stress exposure in mice. Increased expression <strong>of</strong> Cyp17a1 gives hints<br />

for hepatic induction <strong>of</strong> steroidogenesis (Akhtar et al. 2005) in repeatedly stressed mice, and upregulation<br />

<strong>of</strong> expression <strong>of</strong> Cyp39a1 and Cyp2b10 indicates increased removal <strong>of</strong> steroids <strong>by</strong> bile<br />

acid production (Li-Hawkins et al. 2000). Importantly, the cholesterol metabolism <strong>of</strong> rodents and<br />

humans is difficult to compare. In mice, HDL is the main lipoprotein present in the blood and<br />

essentially required for steroidogenesis (Martin et al. 1999, Ricart-Jané et al. 2002), whereas<br />

humans use LDL for steroid synthesis and have lower HDL concentrations (Chen W et al. 2005,<br />

Martin et al. 1999). Furthermore, HDL is able to scavenge endotoxins from the plasma (Barter et<br />

al. 2004). The relevance <strong>of</strong> stress-induced elevation <strong>of</strong> plasma HDL levels in mice, e. g. for steroid<br />

synthesis or scavenging the bacterial compound lipopolysaccharide, remains to be elucidated.<br />

Other authors found that lipopolysaccharide or TNF challenges particularly increase the catabolic<br />

rate (Ogimoto et al. 2006, Sugita et al. 2002). Such effects seem to be mediated via TNF-induced<br />

neuroendocrine stimulation, e. g. activation <strong>of</strong> the sympathetic nervous system that primarily<br />

causes glucose formation (Bag<strong>by</strong> et al. 1992, Finck et al. 1998, Finck/Johnson 2000,<br />

173


Maren Depke<br />

Discussion and Conclusions<br />

Leibowitz/Wortley 2004, Lundberg 2005, McGuinness et al. 1999, Ogimoto et al. 2006, Sugita et<br />

al. 2002). In addition, release <strong>of</strong> catecholamines potentially induces the expression <strong>of</strong> the leptin<br />

gene in adipose and hepatic tissue (Arnalich et al. 1999, Finck/Johnson 2000, Mak et al. 2006).<br />

Leptin then can signal afferently to the brain to sustain inhibitory effects on food intake. In this<br />

relation, one should expect reduced food consumption in the hyperleptinemic repeatedly<br />

stressed mice. However, food intake was not altered, whereas total body mass furthermore was<br />

lost. One possible explanation for the drastic loss <strong>of</strong> body weight along with hyperleptinemia is<br />

that leptin potentially can increase the energy expenditure such as enhancing the activation <strong>of</strong><br />

the respiratory chain and, therefore, can increase energy consumption (Fleury et al. 1997,<br />

Ricquier/Bouillaud 2000). In this study, already acute stress drastically induced changes <strong>of</strong><br />

hepatic gene expression that did not significantly disrupt allostatic regulation <strong>of</strong> metabolism in<br />

mice. In turn, repeated psychological stress in BALB/c mice along with systemic<br />

immunodeficiency that was reported previously (Kiank et al. 2006, 2007b) induced a<br />

hypermetabolic stress syndrome.<br />

It is not clear why some individuals lose weight during prolonged stressful situations, whereas<br />

others gain body mass (Alberda et al. 2006, Harris et al. 1998, Morley et al. 2006,<br />

Vanhorebeek/Van den Berghe 2004, Wilmore 2000, Wray et al. 2002). Because there is an<br />

increased number <strong>of</strong> patients suffering from metabolic syndrome and clinically relevant<br />

associated illness, many publications show that chronic stress is promoting the development <strong>of</strong> a<br />

metabolic syndrome that is associated with gain <strong>of</strong> fat mass (obesity), type 2 diabetes,<br />

hyperlipidemia, and hypertension (Alberda et al. 2006, Harris et al. 1998, Lundberg 2005). In<br />

contrast, the animal experiments with BALB/c mice as a mouse strain with high stress<br />

susceptibility (Kiank et al. 2006) led to the detection <strong>of</strong> metabolically driven wasting because <strong>of</strong> a<br />

hypercatabolic stress response. It can be assumed that the genetic predisposition influences the<br />

development <strong>of</strong> either stress-induced metabolic syndrome or loss <strong>of</strong> body weight phenotype.<br />

Moreover, it is shown that besides genetic predisposition, environmental factors influence<br />

prenatal and postnatal neuronal and neuroendocrine differentiation, resulting in different coping<br />

styles in the adult (Koolhaas et al. 2007). They showed that proactive/aggressive animals<br />

developed stress-induced hypertension, cardiac arrhythmia, and inflammation, whereas<br />

reactive/passive individuals are more susceptible to anxiety disorders, metabolic syndrome,<br />

depression, and infection. However, the neurobiology and endocrine regulation <strong>of</strong> these different<br />

coping styles are not well understood yet. A loss <strong>of</strong> biological reserves as in the model <strong>of</strong><br />

repeated stress exposure is as fatal as the development <strong>of</strong> a metabolic syndrome because <strong>of</strong><br />

losing the ability to fight infection and cancer (Alberda et al. 2006, Hang et al. 2003, Hansen MB<br />

et al. 1998, Morley et al. 2006, Souba et al. 1985).<br />

In the clinical setting, it is now clear that the catabolic response will become autodestructive if<br />

not contained. The severity <strong>of</strong> complications will occur in proportion to lost body protein. In the<br />

model <strong>of</strong> repeatedly stressed mice, particularly arginine deficiency became evident. Interestingly,<br />

alimentation with arginine and omega-3 fatty acids-enriched enteral feeds decreased hospital<br />

days and infectious complications in critically ill patients (Beale et al. 1999), which commonly<br />

show a loss <strong>of</strong> about 10 % <strong>of</strong> lean body mass (Arnalich et al. 1999, Fleury et al. 1997,<br />

Ricquier/Bouillaud 2000). Healthy adults require about 0.8 g protein/kg body weight·d to<br />

maintain homeostasis. Stressful events such as traumatic injury or infection increase the body’s<br />

protein requirement up to 1.5–2 g protein/kg body weight·d or even more. However, humans<br />

cannot metabolize more than 2 g/kg body weight·d. This <strong>of</strong>ten results in a fatal negative nitrogen<br />

174


Maren Depke<br />

Discussion and Conclusions<br />

balance in severely ill patients (Demling 2007). Finally, amplified protein breakdown with a loss <strong>of</strong><br />

more than 40 % <strong>of</strong> lean body mass leads to irreversible cell damage (Beale et al. 1999, Demling<br />

2007).<br />

In conclusion, the physiological measurements and hepatic gene expression pr<strong>of</strong>iling<br />

demonstrated that highly demanding psychological stress in the absence <strong>of</strong> injury or infection is<br />

able to induce a severe hypermetabolic syndrome in mice. Such overwhelming wasting condition<br />

can reduce the individual’s resistance to further stressful stimuli such as injury or infection.<br />

The same data set <strong>of</strong> hepatic gene expression was used to gain deeper insight into hepatic<br />

immune regulatory and cell cycle processes <strong>of</strong> BALB/c mice, which develop systemic<br />

immunodeficiency with a reduced antibacterial defense after repeated stress exposure (Kiank et<br />

al. 2006, 2007a,b). This second part <strong>of</strong> the results has been published <strong>by</strong> Depke et al. in 2009.<br />

Already a single acute stress exposure drastically altered the expression <strong>of</strong> immune response<br />

and <strong>of</strong> apoptosis related genes. An activation <strong>of</strong> immunosuppressive mechanisms in the liver was<br />

measured in both acutely and chronically stressed mice including increased expression <strong>of</strong><br />

Tsc22d3 (GILZ, glucocorticoid-induced leucine zipper) and Fkbp5, which both are inducible <strong>by</strong><br />

high glucocorticoid levels (Ayroldi et al. 2007, Berrebi et al. 2003, D'Adamio et al. 1997,<br />

Mittelstadt/Ashwell 2001). Tsc22d3 inhibits the expression <strong>of</strong> CD80 and CD86 co-stimulatory<br />

molecules or toll-like receptor 2 and reduces the production <strong>of</strong> proinflammatory mediators<br />

(Berrebi et al. 2003, Cohen et al. 2006). Fkbp5, an immunophilin family member, inhibits<br />

calcineurin signaling in lymphocytes and modifies glucocorticoid signaling <strong>by</strong> binding heat shock<br />

proteins <strong>of</strong> steroid receptors (Baughman et al. 1995, Sinars et al. 2003). Other B and T<br />

lymphocyte signaling molecules such as protein kinase C ν (Prkcn), or several GTPases showed<br />

altered hepatic mRNA expression in both acutely and chronically stressed mice, respectively.<br />

Importantly, the down-regulation <strong>of</strong> several IFN-γ inducible genes after acute stress was<br />

amplified <strong>by</strong> repeated stress exposure. Some <strong>of</strong> the interferon-inducible mRNA transcripts like<br />

Igtp and Stat1, which were repressed in the liver <strong>of</strong> chronically stressed mice, encode for <strong>host</strong><br />

proteins <strong>of</strong> well characterized antimicrobial activity (Döffinger et al. 2002, Johnson/Scott 2007,<br />

Martens et al. 2004). Diminished expression <strong>of</strong> IFN-γ inducible GTPases was shown to be related<br />

to a reduced defense against Listeria monocytogenes and Mycobacterium avium infection<br />

(Martens et al. 2004). Stat1 is important for IFN-γ, IL-6, IL-10, and IL-12 signaling and for<br />

maintaining MHC- and co-stimulatory molecule expression in dendritic cells, which all are<br />

important for the antibacterial response (Döffinger et al. 2002, Johnson/Scott 2007). In line with<br />

this, chronically stressed animals showed a repression <strong>of</strong> the Stat1 gene and reduced expression<br />

<strong>of</strong> MHC-molecules along with a heightened susceptibility to bacterial infections (Kiank et al. 2006,<br />

2007b) and an increased spreading <strong>of</strong> translocated commensals into liver and lung, which went<br />

along with a reduced ex vivo inducibility <strong>of</strong> IFN-γ (Kiank et al. 2008).<br />

Besides the reduced expression <strong>of</strong> cellular signaling molecules, genes associated with immune<br />

cell migration and tissue invasion were differentially expressed. These included T cell<br />

chemoattractive peptides such as Cxcl11 and Cxcl12 and neutrophil chemoattractive Cxcl1<br />

(Ghosh et al. 2006, Helbig et al. 2004, Wiekowski et al. 2001) and an increased expression <strong>of</strong> cell<br />

adhesion molecules such as Vcam-1 in the liver homogenate after both acute and chronic stress.<br />

Arhgap5 was selectively induced after repeated stress exposure (Cook-Mills 2002, Haddad/Harb<br />

2005, Kim et al. 2006, Petri/Bixel 2006) while the expression <strong>of</strong> several claudins, which are tight<br />

junction proteins that regulate paracellular permeability (Amasheh et al. 2002), was significantly<br />

175


Maren Depke<br />

Discussion and Conclusions<br />

reduced. Thus, data from the murine psychological stress model indicate that leukocyte<br />

migration associated molecules were increasingly transcribed already after acute stress.<br />

However, leukocyte infiltration into hepatic tissue was not measurable at this particular moment<br />

immediately after a single stress exposure but this may be due to the point in time <strong>of</strong> data<br />

acquisition which was chosen. Other authors showed leukocyte recruitment into the liver 3−6 h<br />

after hepatic ischemia/reperfusion (Zwacka et al. 1997) or 2−4 h after initiation <strong>of</strong> an acute phase<br />

response (APR) following brain injury (Campbell et al. 2003). Therefore, one can propose that cell<br />

recruitment into the liver may be detectable in the next few hours after termination <strong>of</strong> acute<br />

stress, which remains to be elucidated in further experiments, and later manifests during<br />

repeated stress exposure as shown <strong>by</strong> the data <strong>of</strong> chronically stressed mice.<br />

An invasion <strong>of</strong> inflammatory cells such as neutrophils and macrophages into tissue is <strong>of</strong>ten<br />

associated with increased release <strong>of</strong> proinflammatory cytokines, which in turn enhance<br />

catabolism and stimulate the hypothalamus-pituitary-adrenal axis (Bartolomucci 2007, Herold et<br />

al. 2006, Lundberg 2005). Although increased expression <strong>of</strong> cytokine genes in the liver was not<br />

shown, plasma glucocorticoids levels were increased after both acute and chronic psychological<br />

stress exposure causing up-regulation <strong>of</strong> glucocorticoid-inducible genes, such as acute phase<br />

proteins (Kiank et al. 2006, Depke et al. 2008). In line with these data after chronic stress, Yoo<br />

and Desiderio found increased expression <strong>of</strong> several APR markers, including C-reactive protein<br />

(CRP), serum amyloid A proteins, lipocalins or orsomucoid at 3 h, 6 h and 12 h after low-dose<br />

bolus application <strong>of</strong> LPS (Yoo/Desiderio 2003), which can serve as a model <strong>of</strong> minute bacterial<br />

translocation. Acute phase proteins (APPs) have several immune and metabolic regulatory<br />

functions such as enhancing the antimicrobial response: CRP can opsonize microorganisms and<br />

activate the complement system (Casey et al. 2008, Mold/Du Clos 2006), apolipoproteins, or<br />

lipocalin 2 regulate the cholesterol transport, endotoxin scavenging, and free radical production<br />

(Levels et al. 2003, Roudkenar et al. 2007, Tseng et al. 2004, Wurfel et al. 1994). In contrast, other<br />

APPs such as haptoglobin, which were found to be induced after repeated stress, have strong<br />

anti-inflammatory effects (Tseng et al. 2004) and therewith may contribute to the reduced<br />

antibacterial defense <strong>of</strong> chronically stressed mice, which even suffered from long-lasting bacterial<br />

infiltration into liver and lung (Kiank et al. 2008). Along with the increased expression <strong>of</strong> APR<br />

genes, high expression levels <strong>of</strong> cytochrome P450 enzymes were detected which normally are<br />

suppressed during an APR (Siewert et al. 2000). This can be explained <strong>by</strong> findings <strong>of</strong> others who<br />

showed that hepatic levels <strong>of</strong> P450 enzymes increased during fasting and then became resistant<br />

to the suppression during inflammatory states (Iber et al. 2001, Morgan 2001). Therefore, the<br />

increased cytochrome expression in the chronic stress model may be a part <strong>of</strong> the hypercatabolic<br />

stress response that overcomes an inflammation-induced repression <strong>of</strong> P450 enzyme expression.<br />

An activation <strong>of</strong> cytochromes enhances the generation <strong>of</strong> reactive oxygen species (ROS) (Morgan<br />

2001, Zangar et al. 2004). Oxidative stress, in turn, can cause lipid peroxidation or increase<br />

protein carbonyl content (Ermak/Davies 2001, Garg/Aggarwal 2002, Ott et al. 2007, Zangar et al.<br />

2004), which was also observed in plasma and liver after acute stress exposure in this study.<br />

Carbonylated proteins are likely non-functional and prone to degradation. Oxidant-induced<br />

damage in hepatic tissue <strong>of</strong> acutely stressed mice is supported <strong>by</strong> the finding that several cell<br />

cycle and cell death associated genes such as Cdkn1a, Gadd45b, Gadd45g or Fkbp5 were already<br />

induced immediately after acute stress – a moment when detection <strong>of</strong> cellular apoptosis is<br />

probably too early – and remained highly expressed after repeated stress when increased<br />

hepatocyte apoptosis was measured.<br />

176


Maren Depke<br />

Discussion and Conclusions<br />

However, tissue protective mechanisms seem to be active in the liver <strong>of</strong> chronically stressed<br />

mice as well. These include an increased expression <strong>of</strong> Alas1, which functions as an oxygen<br />

radical scavenger (Kaliman et al. 2005), and <strong>of</strong> glutaredoxin and glutathione S-transferase alpha<br />

that are ubiquitous proteins with redox-active cysteine-groups (Garg/Aggarwal 2002,<br />

Haddad/Harb 2005, Jurado et al. 2003). Increased expression <strong>of</strong> these anti-oxidant enzymes may<br />

compensate and protect from further oxidative stress-induced damage. Therefore, long lasting<br />

increased glucocorticoid levels, which are highly potent mediators <strong>of</strong> inducing apoptotic<br />

processes, are proposed to be the main mediators <strong>of</strong> hepatic cell death in chronically stressed<br />

animals (Ayroldi et al. 2007, Herold et al. 2006), but this remains to be elucidated.<br />

The liver has an extraordinary ability to regenerate or heal itself <strong>by</strong> replacing or repairing<br />

injured tissue (Fausto et al. 2006, Klein et al. 2005, Riehle et al. 2008). Hepatic gene expression<br />

pr<strong>of</strong>iling after acute and chronic stress included increased hepatic expression <strong>of</strong> protein tyrosine<br />

phosphatase 4a1, which initiates the liver regeneration response, and <strong>of</strong> IL-6 receptor whose<br />

ligation was shown to be protective during liver injury (Fausto et al. 2006, Klein et al. 2005, Riehle<br />

et al. 2008). Thus, acute stress-induced up-regulation <strong>of</strong> genes coding for tissue destructive<br />

mechanisms associated with oxidative stress were counter-regulated <strong>by</strong> increased expression <strong>of</strong><br />

anti-oxidative and tissue regenerative genes which may have protected mice from more severe<br />

damage <strong>of</strong> hepatic tissue during chronic psychological stress.<br />

In summary, the data show that already acute stress exposure induces hepatic mRNA<br />

expression <strong>of</strong> several LPS and glucocorticoid-sensitive immune response genes as well as <strong>of</strong><br />

apoptosis-related markers. Besides increased oxidative stress, this does not cause measurable<br />

immune alterations or cell death immediately after stress exposure. However, when stress is<br />

repeated and becomes chronic, reduced antigen presentation and altered gene expression <strong>of</strong><br />

cellular signaling molecules <strong>of</strong> immune cells along with an ongoing expression <strong>of</strong> apoptosisrelated<br />

molecules may contribute to biologically relevant immunosuppression which was<br />

demonstrated <strong>by</strong> a reduced ability <strong>of</strong> stressed mice to clear bacterial infections from the liver.<br />

Moreover, the induction <strong>of</strong> cell protective and liver regenerative genes in hepatic tissue <strong>of</strong><br />

chronically stressed mice give strong evidence that counter-regulatory mechanisms are activated<br />

to prevent further hepatocyte damage in these animals.<br />

177


Maren Depke<br />

Discussion and Conclusions<br />

KIDNEY GENE EXPRESSION PATTERN IN AN<br />

IN VIVO INFECTION MODEL<br />

The SigB regulon contains several virulence associated genes, and expression <strong>of</strong> this regulon in<br />

general promotes adhesion and reduces production <strong>of</strong> extracellular toxins (Bisch<strong>of</strong>f et al. 2004).<br />

While regulation <strong>of</strong> SigB activity has already been studied in vitro (Senn et al. 2005), its impact on<br />

in vivo physiology, e. g. in infection models, is still matter <strong>of</strong> debate. The aim <strong>of</strong> this study was the<br />

analysis <strong>of</strong> the relevance <strong>of</strong> SigB-dependent gene expression regulation for the adaptation <strong>of</strong> the<br />

<strong>host</strong> transcriptome in murine infection.<br />

In an i. v. murine infection model the two Staphylococcus aureus strains RN1HG and its<br />

isogenic sigB mutant were used to infect mice with comparable infection doses. In mouse kidney,<br />

the infection resulted in similar infection rates <strong>of</strong> approximately 1.0E+06 cfu/10 mg <strong>of</strong> tissue. The<br />

role <strong>of</strong> SigB has already been analyzed in different in vivo infection models. A general observation<br />

<strong>of</strong> these studies was that sigB deletion mutants and their parental strain can accomplish<br />

comparable bacterial loads. In the clinical isolate WCUH29 the allelic replacement <strong>of</strong> sigB did not<br />

lead to differing bacterial load in organs or wounds after testing original strain and isogenic<br />

mutant in different types <strong>of</strong> infection models (Nicholas et al. 1999). Also Chan et al. (1998)<br />

observed similar numbers <strong>of</strong> cfu in a skin abcess model, but the results might have been<br />

compromised <strong>by</strong> the fact that the group used a sigB deletion in the 8325-4 background which<br />

itself harbors an inactivating mutation in rsbU, a gene for a SigB regulatory phosphatase<br />

(Kullik/Giachino 1997). Horsburgh et al. published in 2002 the construction <strong>of</strong> S. aureus SH1000, a<br />

variant <strong>of</strong> strain 8325-4, in which the rsbU gene had been repaired. SH1000 and 8325-4 resulted<br />

in the recoverage <strong>of</strong> similar numbers <strong>of</strong> cfu in a murine subcutaneous skin abcess model<br />

(Horsburgh et al. 2002b). Contrarily, another study reported a significantly different bacterial<br />

burden in kidney after i. v. infection with S. aureus 8325-4 (RsbU − ), SH1000 (RsbU + ), and a sigB<br />

knockout mutant <strong>of</strong> SH1000 (SigB − ). In this setting, the repair <strong>of</strong> rsbU leads to an increased cfu<br />

count 14 days after inoculation (Jonsson et al. 2004). In the study described in this thesis, almost<br />

equal mean values <strong>of</strong> infection rate were observed for S. aureus RN1HG and its isogenic sigB<br />

mutant, a further confirmation <strong>of</strong> literature data indicating similar virulence <strong>of</strong> sigB deficient and<br />

positive S. aureus strains.<br />

In order to gain a much more detailed view <strong>of</strong> the <strong>host</strong> reaction to both strains, kidney gene<br />

expression pr<strong>of</strong>iles were compared using Affymetrix GeneChip arrays. In a total number <strong>of</strong> 19<br />

infected samples, highly similar gene expression pr<strong>of</strong>iles were recorded 4 or 5 days after infection<br />

with RN1HG or its sigB derivative. Even with this sophisticated analysis no difference in the<br />

expression pattern in murine kidney after infection with the two strains was observed, indicating<br />

that SigB indeed does not influence the <strong>host</strong> response.<br />

However, the comparison to sham-infected samples revealed a rather strong reaction <strong>of</strong> the<br />

murine <strong>host</strong> to infection, bacterial accumulation and proliferation in the kidney. Several immune<br />

response pathways were induced on transcriptional level indicating a vivid activation <strong>of</strong> the<br />

immune response. This applies first to pathways <strong>of</strong> the innate immune system: Pattern<br />

recognition receptors (PRR) like toll-like receptors (TLR) and complement system components<br />

178


Maren Depke<br />

Discussion and Conclusions<br />

were induced and strengthened the first line <strong>of</strong> defense against the <strong>pathogen</strong>. Second, several<br />

immune response signaling pathways exhibited increased gene expression <strong>of</strong> their members, e. g.<br />

proinflammatory pathways <strong>of</strong> IFN, IL-6, and TREM1 signaling, and several proinflammatory<br />

cytokines were also induced, e. g. TNFα, IL-6, RANTES/Ccl5, MCP-1/Ccl2, MCP-3/Ccl7, and MIP-<br />

1α/Ccl3. Members <strong>of</strong> IL-10 signaling, like the IL-10 receptor α-chain (Il10ra) and the signal<br />

transducer STAT3, were induced, too. This pathway aims at limitation <strong>of</strong> the cellular<br />

proinflammatory response. But as the ligand IL-10 was not induced and STAT3 is also part <strong>of</strong> the<br />

IL-6 signaling, it is not clear whether the IL-10 signaling was active at the time point <strong>of</strong> analysis or<br />

whether it was only prepared to receive signals in a later phase <strong>of</strong> infection. As third example, the<br />

infected samples exhibited a distinct transcriptional induction <strong>of</strong> several members <strong>of</strong> the<br />

presentation pathways for intracellular and extracellular antigens via MHC molecules. The<br />

comparison <strong>of</strong> infection and sham infection additionally revealed metabolic disturbances shifting<br />

to both anabolic and catabolic direction <strong>of</strong> metabolism. This might indicate a high extent <strong>of</strong><br />

infection and illness that impairs organ function, nutrition supply and possibly oxygen availability.<br />

The comparison <strong>of</strong> S. aureus infected samples with non-infected controls proved the strong<br />

reaction <strong>of</strong> the <strong>host</strong> to infection. But in the model described in this thesis the <strong>host</strong> reaction does<br />

not differ depending on the strain used for infection.<br />

Until now, the role <strong>of</strong> sigB in infections has not been completely clarified. There is evidence<br />

for an influence <strong>of</strong> sigB on bi<strong>of</strong>ilm formation. Attachment and microcolony formation are the two<br />

central steps in the early bi<strong>of</strong>ilm formation. In an in vitro bi<strong>of</strong>ilm formation model (using<br />

polystyrene microtiter plates) it was shown that increased sigB expression leads to increased<br />

attachment and to increased microcolony formation with increased size <strong>of</strong> staphylococcal<br />

autoaggregates (Bateman et al. 2001).<br />

In detail, bi<strong>of</strong>ilm formation is mediated <strong>by</strong> polysaccharide intercellular adhesin/PIA, an<br />

extracellular adhesin <strong>of</strong> staphylococci composed <strong>of</strong> poly-N-acetylglucosamine/PNAG. Although<br />

highly conserved in staphylococcal strains, the ica operon, coding for enzymes responsible for PIA<br />

synthesis, is expressed in vitro only in a few strains. For the mucosal isolate MA12 it was proven<br />

that a sigB insertion mutant (sigB::ermB) lost the bi<strong>of</strong>ilm formation ability after osmotic stress<br />

(3 % NaCl) compared to the wild type strain. Additionally, it was demonstrated under osmotic<br />

stress conditions that the mutant exhibited strongly reduced ica expression compared to the wild<br />

type (Rachid et al. 2000).<br />

Even more, a major part <strong>of</strong> ica expression and bi<strong>of</strong>ilm formation was proven to be mediated<br />

<strong>by</strong> SarA. It is known that SigB induces sarA expression, but it has not been demonstrated yet<br />

whether the influence on ica expression is an indirect effect <strong>of</strong> SigB or whether SarA acts<br />

independently <strong>of</strong> SigB in this context. Interestingly, experiments predict the existence <strong>of</strong> a SigBactivated<br />

factor, which either might degrade PIA or repress its synthesis. This factor in turn is<br />

repressed <strong>by</strong> SarA (Valle et al. 2003).<br />

Complementary information regarding the influence <strong>of</strong> sigB knock-out on bi<strong>of</strong>ilm formation<br />

and virulence is available from device-associated staphylococcal infection models in mice. When<br />

MA12 and its isogenic sigB mutant were applied to a central venous catheter (CVC) and later to<br />

the blood stream, both strains were able to form multilayered bi<strong>of</strong>ilms inside the catheter.<br />

Therefore, sigB is not essential for bi<strong>of</strong>ilm formation. On the other hand, there were structural<br />

differences between the bi<strong>of</strong>ilms <strong>of</strong> both strains: The mutant’s bi<strong>of</strong>ilm was lacking certain<br />

extracellular substance. SigB is thus proposed to affect the ability <strong>of</strong> spreading or release from<br />

adhesive sites/bi<strong>of</strong>ilms. In the quantitation <strong>of</strong> total bacterial burden in the inner organs, a<br />

179


Maren Depke<br />

Discussion and Conclusions<br />

significant decrease was apparent after infection with the sigB mutant compared to the parental<br />

strain (Lorenz et al. 2008). Others have published that sigB mutations have no influence on cfu<br />

counts. This emphasizes the important influence <strong>of</strong> other experimental parameters like infection<br />

dose, infection route, site <strong>of</strong> infection or its manifestation, analysis time point after infection, the<br />

<strong>host</strong> factors – at least in parts determined <strong>by</strong> the <strong>host</strong> genome – acting on the bacterium, and the<br />

genetic background <strong>of</strong> the <strong>pathogen</strong>.<br />

Repair <strong>of</strong> rsbU or sigB overexpression in the genetic background <strong>of</strong> BB255, a rsbU mutated<br />

S. aureus strain, leads to an increase in transcription <strong>of</strong> clfA, mediated via a SigB-promoter in the<br />

clfA gene, and an increase in transcription <strong>of</strong> fnbA which is indirectly caused <strong>by</strong> SigB and probably<br />

mediated via a positive effect <strong>of</strong> SigB on sarA and subsequently <strong>of</strong> SarA on the fnbA transcription.<br />

The sigB overexpression mutant features an increased attachment to fibrinogen and fibronectin<br />

coated surfaces and to platelet-fibrin clots, which imitate the environment in injured<br />

endothelium <strong>of</strong> blood vessels. When the same mutant is applied in an infection setup focusing on<br />

adherence <strong>by</strong> inducing catheter-associated aortic vegetation in rat, a transient effect <strong>of</strong> higher<br />

bacterial density in early stages <strong>of</strong> infection was observed. After long-term infection, the<br />

advantage <strong>of</strong> the early phase was counterbalanced in comparison to sigB deficient strains. The<br />

authors argue that S. aureus is still able to produce sufficient amounts <strong>of</strong> adhesins leading to<br />

aortic vegetation even without a functional sigB (Entenza et al. 2005).<br />

In a murine sepsis and arthritis model, the SigB positive strain SH1000 (rsbU repaired) caused<br />

more severe infection than strain 8325-4 which is phenotypically SigB − (11 bp rsbU deletion) as<br />

illustrated <strong>by</strong> higher mortality, more pronounced weight loss and higher serum levels <strong>of</strong> IL-6 as<br />

indicator <strong>of</strong> inflammation. Additionally, at day 14 after i. v. infection higher arthritis frequency<br />

and severity were observed after infection with SH1000. In an attempt to elucidate whether the<br />

effect originated from increased elimination <strong>of</strong> the SigB negative strain or a decreased virulence<br />

in situ the authors infected mice intra-articularly. In this experiment no difference in the ability to<br />

cause inflammation was recognized between infections with the two strains. This led to the<br />

conclusion that SigB-influenced steps before arthritical joint involvement contributed to the<br />

differences between the strains after i. v. infection, possibly including SigB regulated cell surface<br />

adhesins as crucial virulence factors (Jonsson et al. 2004).<br />

In summary, the impact <strong>of</strong> SigB in vivo is still disputed, and results in literature references are<br />

partly inconsistent. A hypothesis reconfirmed <strong>by</strong> several authors is that the effect <strong>of</strong> SigB in<br />

in vivo infection models possibly only manifests in early phases <strong>of</strong> the infection and only impacts<br />

processes in a transient manner.<br />

Hence, it can be speculated that in the model described in this thesis the analyzed time point<br />

might have been too late for detection <strong>of</strong> differences in the <strong>host</strong> reaction.<br />

Furthermore, the effects <strong>of</strong> SigB are part <strong>of</strong> a complex and interfering regulation mechanism<br />

including several other regulatory molecules like RNAIII from the agr locus or SarA. SigB<br />

dependent transcription <strong>of</strong> sar locus was demonstrated (Ziebandt et al. 2004). The SarA protein is<br />

a transcriptional regulator, which represses (e. g. spa) or activates (e. g. fnbA) gene expression<br />

directly. It also activates expression <strong>of</strong> the agr system and therefore influences indirectly the<br />

expression <strong>of</strong> genes responding to RNAIII (Chien et al. 1999). SigB itself has a negative effect on<br />

the amount <strong>of</strong> RNAIII (Horsburgh et al. 2002b). Several genes or their promoter regions possess<br />

binding sites for more than one transcriptional regulator or sigma factor, e. g. the sae locus coded<br />

two-component system also influences SigB-regulated target genes (Goerke et al. 2005).<br />

Therefore, the transcriptional pattern <strong>of</strong> the regulators overlap, and the mutation <strong>of</strong> one <strong>of</strong> them<br />

180


Maren Depke<br />

Discussion and Conclusions<br />

might be compensated at least in parts <strong>by</strong> others. Such compensation might contribute to the<br />

non-distinguishable <strong>host</strong> reaction in the model described in this thesis. Additionally, the balance<br />

between the different regulators might be different in in vivo settings than in the well-studied<br />

in vitro conditions. Goerke et al. (2005) observed less SigB activity in vivo in a guinea pig model <strong>of</strong><br />

implant infection than in vitro under induced conditions like in the post-exponential growth<br />

phase. But differences were also observed between in vitro and in vivo analyses in the sequential<br />

order <strong>of</strong> gene expression pattern (Goerke et al. 2005).<br />

Even more complicated, a transposon mutagenesis study identified 13 SigB-independently<br />

transcribed genes whose mutation led to an increase in hla transcription and protease activity.<br />

This effect had a different magnitude depending <strong>of</strong> the mutation-affected gene and could be<br />

traced back to a gradual reduction <strong>of</strong> SigB activity after mutation. As these genes are part <strong>of</strong><br />

normal cellular metabolic pathways and do not exhibit DNA binding motifs the authors propose<br />

them to be indirectly involved in the control <strong>of</strong> RsbU activity, preventing full signaling through<br />

RsbU but not fully blocking the phosphatase activity. Such hypothesis assigns these metabolic<br />

genes a function similar to that <strong>of</strong> the B. subtilis RsbRSTX regulation system sensing<br />

environmental conditions, which is absent in S. aureus (Shaw et al. 2006).<br />

In this context, the question arises whether activation <strong>of</strong> SigB really takes place in the specific<br />

in vivo infection setting or whether there are differences depending on the model used, the site<br />

<strong>of</strong> infection or other experimental factors influencing S. aureus during infection. Reduced in vivo<br />

activity <strong>of</strong> SigB might also result in the similarity <strong>of</strong> <strong>host</strong> reaction to infection with S. aureus<br />

RN1HG and its sigB mutant as described in this thesis. The question whether sigB and the SigB<br />

regulon are really expressed in infection models strongly suggests the examination <strong>of</strong> sigB and<br />

SigB-dependent marker genes <strong>by</strong> real-time qPCR in infected samples for further investigations on<br />

the relevance <strong>of</strong> SigB in in vivo settings.<br />

In summary, the results <strong>of</strong> this study do not provide any hints for differences in the<br />

<strong>pathogen</strong>esis or pathomechanism <strong>of</strong> the S. aureus strains RN1HG and ΔsigB in the selected model<br />

<strong>of</strong> i. v. infection in mice. If really existing, such differences might be transient and only apparent<br />

at earlier time points. Effects <strong>of</strong> SigB might also be superimposed in in vivo infection <strong>by</strong> the<br />

interlaced pattern <strong>of</strong> other regulators. There is also the possibility <strong>of</strong> missing activity <strong>of</strong> SigB<br />

in vivo which could explain the similarity <strong>of</strong> <strong>host</strong> reaction to infection with S. aureus RN1HG and<br />

its sigB mutant in the model used in this study. SigB might possess only to a lesser extent<br />

characteristics attributed to virulence factors and might act in vivo more like a virulence<br />

modulator and fine tune bacterial reactions. Assuming such function, the missing <strong>of</strong> detectable<br />

differences in the <strong>host</strong>’s reaction to S. aureus RN1HG and its isogenic sigB mutant is explainable.<br />

181


Maren Depke<br />

Discussion and Conclusions<br />

GENE EXPRESSION PATTERN OF BONE-MARROW DERIVED<br />

MACROPHAGES AFTER INTERFERON-GAMMA TREATMENT<br />

The phagocytes <strong>of</strong> the innate immune system, neutrophils, monocytes/macrophages and<br />

dendritic cells, accomplish central functions during the encounter <strong>of</strong> <strong>host</strong> and <strong>pathogen</strong>. They are<br />

the first group <strong>of</strong> immune cells which recognize and react to an infection and the main effectors<br />

<strong>of</strong> clearance <strong>of</strong> <strong>pathogen</strong>s. Beside the different location in the non-reactive situation, when<br />

neutrophils and monocytes circulate in the blood stream and macrophages and dendritic cells are<br />

found in the tissue, macrophages and dendritic cells are long-living in comparison to the shortliving<br />

neutrophils. Macrophages and dendritic cells have important functions in the regulation <strong>of</strong><br />

the immune reaction, e. g. <strong>by</strong> their antigen presentation ability (Gordon S 2007). The central<br />

position <strong>of</strong> macrophages in the immune defense accounts for the relevance <strong>of</strong> research on<br />

macrophage related topics.<br />

Accordingly, the study described in this thesis aimed to analyze on a molecular level the<br />

reaction <strong>of</strong> macrophages to interferon-γ (IFN-γ), which in low doses is known to be a priming<br />

signal for macrophages and prepares the cells for a faster reaction to a second stimulus (Dalton<br />

et al. 1993, Huang et al. 1993, Ma J et al. 2003, Mosser 2003). Here, <strong>by</strong> utilization <strong>of</strong> bonemarrow<br />

derived macrophages (BMM), which were differentiated in vitro from bone marrow stem<br />

cells under the influence <strong>of</strong> granulocyte-macrophage colony stimulating factor (GM-CSF), any<br />

influence <strong>of</strong> immunological conditioning or in vivo stimuli were avoided. Such influences could<br />

impair the experimental results when using mature macrophages prepared from animal organs.<br />

Furthermore, recently established serum-free culture conditions (Eske et al. 2009) were applied.<br />

This new system circumvented uncontrollable influences on BMM experiments, which would<br />

have been introduced <strong>by</strong> vendor- and batch-varying, cytokine-, hormone- or endotoxincontaining<br />

serum if traditional cultivation conditions had been applied.<br />

The study included BMM <strong>of</strong> the two mouse strains BALB/c and C57BL/6, which were chosen<br />

because <strong>of</strong> the differences observed between these mouse strains in in vivo and in vitro infection<br />

studies (Breitbach et al. 2006, Autenrieth et al. 1994, van Erp et al. 2006). Thus, the data set from<br />

this study was used to compare on the one hand non-stimulated control BMM <strong>of</strong> both strains,<br />

and on the other hand BMM <strong>of</strong> the two strains after IFN-γ treatment.<br />

Another feature <strong>of</strong> this study was the combined approach <strong>of</strong> transcriptome (Maren Depke)<br />

and proteome (Dinh Hoang Dang Khoa) analysis. In the comparison <strong>of</strong> transcriptome and<br />

proteome results, it was expected that the microarray as whole genome array covered a much<br />

higher number <strong>of</strong> genes than the number <strong>of</strong> proteins covered <strong>by</strong> the gel-free LC-MS/MS<br />

proteome approach, which was already chosen as best comparable method. Anyway, a defined<br />

part <strong>of</strong> the microarray results like information on genes encoding membrane or secreted proteins<br />

can practically not match proteome results for lack <strong>of</strong> accessibility. Hence, it was not surprising to<br />

find a high number <strong>of</strong> differentially expressed genes, for which protein data were not available.<br />

Vice versa, corresponding gene expression values were recorded for almost all protein data.<br />

A good agreement <strong>of</strong> the results from both analysis levels was evident for the IFN-γ effects.<br />

Depending on the selected comparison, the overlap <strong>of</strong> differentially expressed genes and<br />

182


Maren Depke<br />

Discussion and Conclusions<br />

proteins different in abundance amounted to approximately 40 % to 60 % in reference to<br />

genes/proteins which were accessible with both methods. Most recently, a study on LPSactivated<br />

RAW 264.7 macrophages and C57BL/6 BMM was published. Applying isotope coded<br />

affinity tagging (ICAT), multidimensional liquid chromatography, and mass spectrometry, 1064<br />

proteins were identified and quantified <strong>of</strong> which 36 differed in abundance between LPS-treated<br />

and control cells. Comparison to microarray data revealed approximately 75% concordance<br />

(Swearingen et al. 2010). Although the treatments in both studies were different, the total,<br />

regulated and overlap numbers were roughly comparable.<br />

Nevertheless, a different phenomenon was observed when analyzing differences between<br />

BMM <strong>of</strong> both strains. Expectedly, a high number <strong>of</strong> differentially expressed genes was observed,<br />

which were not accessible with the proteome approach. But also a high number <strong>of</strong> regulated<br />

proteins was recorded, for which transcriptome data were mostly available but did not show<br />

differential expression. Thus, BMM in both strains might differ in post-transcriptional,<br />

translational, and protein stability/turnover processes.<br />

Although the regulated genes/proteins were not always identical for the different<br />

comparisons, similar global results were received <strong>by</strong> both approaches.<br />

1) IFN-γ treatment mainly led to induction <strong>of</strong> gene expression or an increase in protein<br />

abundance.<br />

2) IFN-γ induced changes were highly similar in BALB/c BMM and C57BL/6 BMM, since many <strong>of</strong><br />

them were observed in BMM <strong>of</strong> both strains or, when observed only in one <strong>of</strong> them, showed a<br />

highly similar trend in the other in reference to their fold change values.<br />

3) In the comparison between BMM <strong>of</strong> the two analyzed mouse strains, about 50 % <strong>of</strong> regulated<br />

genes/proteins exhibited higher expression/abundance in BALB/c BMM whereas the other<br />

50 % showed higher expression/abundance in C57BL/6 BMM.<br />

4) Strain differences were highly similar in the comparison <strong>of</strong> control level BMM and in the<br />

comparison <strong>of</strong> IFN-γ treated BMM.<br />

Apart from the comparison to proteome results, the transcriptome data were analyzed for<br />

their biological content. The confirmation <strong>of</strong> known IFN-γ effects in the data set proved the<br />

relevance <strong>of</strong> the results. Very prominent, the induction <strong>of</strong> immunoproteasome and antigen<br />

presentation genes became visible. These included Psmb8, Psmb9, Psmb10, Psme1, and Psme2,<br />

genes <strong>of</strong> proteasome subunits, Tap1 and Tap2, peptide transporters from cytosol to the<br />

endoplasmatic reticulum (ER), Erap1, ER aminopeptidase, and MHC class I and class II genes,<br />

which were already described to be induced <strong>by</strong> interferon (Aki et al. 1994, Van den Eynde/Morel<br />

2001, Brucet et al. 2004, Ma W et al. 1997, Schiffer et al. 2002, Chang et al. 2005, Jung et al.<br />

2009, Benoist/Mathis 1990, Boehm et al. 1997, Gobin/van den Elsen 2000). IFN-γ is also in vivo<br />

responsible for the complete exchange <strong>of</strong> the constitutive to the immunoproteasome since this<br />

exchange was shown to be reduced to 50 % in the liver <strong>of</strong> IFN-γ -/- BALB/c mice after infection with<br />

lymphocytic choriomeningitis virus. The authors suppose TNF-α to be responsible for the<br />

remaining fraction <strong>of</strong> exchange (Khan S et al. 2001). Other in vivo experiments using the fungal<br />

<strong>pathogen</strong> Histoplasma capsulatum in infections <strong>of</strong> IFN-γ -/- C57BL/6 mice resulted in a complete<br />

inability to induce the immunoproteasome. In that study, the authors could not find hints that<br />

other cytokines can compensate for the loss <strong>of</strong> IFN-γ in reference to the immunoproteasome<br />

(Barton et al. 2002). The knockout <strong>of</strong> inducible proteasome β-subunit LMP-7 (Psmb8) led to a<br />

strongly increased susceptibility <strong>of</strong> mice during Toxoplasma gondii infections, which was<br />

183


Maren Depke<br />

Discussion and Conclusions<br />

explained <strong>by</strong> the missing <strong>pathogen</strong>ic antigen processing in dendritic cells and the consequently<br />

missing activation <strong>of</strong> specific cytotoxic CD8 + T cells (Tu et al. 2009).<br />

The induction <strong>of</strong> immunoproteasome β i -subunits but also <strong>of</strong> Tap1 was reported to be<br />

mediated <strong>by</strong> STAT1 during signal transduction and IRF-1, which acts as transcriptional activator<br />

(White et al. 1996, Chatterjee-Kishore et al. 1998, Foss/Prydz 1999, Brucet et al. 2004, Namiki et<br />

al. 2005). In agreement with literature information, IFN-γ treatment led to a clear induction <strong>of</strong><br />

STAT1 and IRF-1 in BMM <strong>of</strong> both strains in the study <strong>of</strong> murine BMM described in this thesis (data<br />

not shown).<br />

The gene expression pr<strong>of</strong>iling revealed the influence <strong>of</strong> IFN-γ on several cytokines. BMM<br />

induced several chemokines (Ccl5, Ccl12, Ccl22, Cxcl9, Cxcl10, Cxcl11, Cxcl16), chemokine<br />

receptors (Ccr5, Ccrl2), interleukins (Il15, Il27), interleukin receptors (Il2rg, Il4ra, Il10ra, Il12rb1,<br />

Il13ra1, Il15ra, Il21r) and also TNF and TNF family ligands (Tnf, Tnfsf10, Tnfsf18) and receptor<br />

(Tnfrsf14). These were interpreted as preparation for a potentially following second stimulus and<br />

consequent immune reaction, which includes pro- and anti-inflammatory directions.<br />

Cxcl9, Cxcl10, and Cxcl11 are known to be induced <strong>by</strong> IFN-γ (Luster et al. 1985, Farber 1990,<br />

Taub et al. 1993, Liao et al. 1995, Cole KE et al. 1998). These chemokines bind the receptor Cxcr3,<br />

which is preferentially found on IL-2 activated Th1 cells (Loetscher M et al. 1996). Of these three<br />

chemokines, Cxcl11 has highest affinity to Cxcr3 and additionally is the most potent agonist<br />

(Cole KE et al. 1998). Because the cytokine exerts its influence only on activated T cells, the<br />

authors additionally infer that the cytokine mainly acts during the immune response and aims to<br />

direct effector T cells to the manifestation site, when T cell activation, IL-2 production, and<br />

proliferation <strong>of</strong> specific T cells had occurred in lymphoid organs (Cole KE et al. 1998).<br />

The three chemokines Cxcl9, Cxcl10, and Cxcl11 exert their function not only in the<br />

chemotaxis <strong>of</strong> Th1 cells via Cxcr3. It was reported that they also can bind to Ccr3, the receptor for<br />

several Ccl chemokines (Ccl5, Ccl7, Ccl8, Ccl11, Ccl13, Ccl24), which is found on Th2 cells. Cxcl9,<br />

Cxcl10, and Cxcl11 act antagonistic when binding to Ccr3. Thus, these chemokines do not only<br />

attract Th1 cells but also inhibit chemotaxis <strong>of</strong> Th2 cells and further polarize the immune<br />

response (Loetscher P et al. 2001). More recently, a similar antagonistic effect <strong>of</strong> Cxcl11 on the<br />

receptor Ccr5 was reported (Petkovic et al. 2004).<br />

Most interestingly, beside agonist and antagonist functions, Cxcl9, Cxcl10, and Cxcl11 have<br />

been reported to exhibit direct antimicrobial functions against Escherichia coli and Listeria<br />

monocytogenes. The observation was rated as biologically relevant since antimicrobially effective<br />

concentrations <strong>of</strong> chemokines were determined in vitro and in vivo (Cole AM et al. 2001). The<br />

antimicrobial effect <strong>of</strong> Cxcl9, Cxcl10, and Cxcl11 also applied to S. aureus (Yang D et al. 2003).<br />

Another study determined constitutive expression <strong>of</strong> Cxcl9 in the human male reproductive<br />

system and secretion into seminal plasma where antibacterial activity against the urogenital<br />

<strong>pathogen</strong> Neisseria gonorrhoeae was demonstrated leading to the conclusion that Cxcl9 is<br />

relevant for the <strong>host</strong>’s local immune defense (Linge et al. 2008). The multifunctional<br />

characteristic <strong>of</strong> chemokine and antimicrobial function is not a unique feature <strong>of</strong> Cxcl9, Cxcl10,<br />

and Cxcl11, but was observed for several other chemokines, too. Antimicrobial activity could be<br />

assigned to about two third <strong>of</strong> all studied chemokines (Yang D et al. 2003).<br />

The cytokine TNF was induced in BMM after IFN-γ treatment. TNF is the central inflammation<br />

mediator. Since long it is known that TNF is the mediator <strong>of</strong> lethal endotoxin effects (Beutler et<br />

al. 1985), which finding was worth a reprint as part <strong>of</strong> the “Pillars <strong>of</strong> Immunology” series in The<br />

Journal <strong>of</strong> Immunology (Vilcek 2008). In the meantime, knowledge was expanded and<br />

184


Maren Depke<br />

Discussion and Conclusions<br />

superfamilies <strong>of</strong> TNF ligands and receptors were identified (Hehlgans/Pfeffer 2005). TNF is<br />

associated with several functions linked to pro-inflammatory effects, which involve a<br />

multiprotein signaling complex at the cell membrane and MAP-kinase signaling and which are<br />

finally mediated via NFκB and c-Jun (Chen G/Goeddel 2002, Wajant et al. 2003). Rarely, TNF can<br />

induce apoptosis via caspase activation, depending on the cellular balance <strong>of</strong> antiapoptosis/proliferation<br />

signals mediated <strong>by</strong> NFκB (Gaur/Aggarwal 2003, Wajant et al. 2003).<br />

Furthermore, TNF is involved in the regulation <strong>of</strong> fever with both pyrogenic and antipyretic<br />

effects depending on physiological or experimental conditions (Leon 2002). In macrophages, TNF<br />

has an important function in controlling intracellular mycobacterial growth (Bekker et al. 2001).<br />

TNF deficient mice are susceptible to Listeria monocytogenes infections, exhibit reduced contact<br />

hypersensitivity and impairment <strong>of</strong> splenic follicular architecture and <strong>of</strong> maturation <strong>of</strong> the<br />

humoral immune response (Pasparakis et al. 1996). Knockout <strong>of</strong> TNF and lymphotoxin-α (TNF-β)<br />

leads to an increased susceptibility to S. aureus infections, but also to a reduced frequency <strong>of</strong><br />

arthritis indicating a damaging influence <strong>of</strong> the cytokine during the inflammatory reaction<br />

(Hultgren et al. 1998). Reflecting the central position <strong>of</strong> TNF in the regulation <strong>of</strong> the immune<br />

response to infection, several strategies <strong>of</strong> <strong>pathogen</strong>s to influence the TNF activity were<br />

reported. The interference can lead to diverse and contrary effects depending on the <strong>pathogen</strong>,<br />

like inhibition <strong>of</strong> apoptosis, enhancement <strong>of</strong> apoptosis, inhibition <strong>of</strong> TNF production or, in case <strong>of</strong><br />

S. aureus, the induction <strong>of</strong> a TNF-like response <strong>by</strong> interaction <strong>of</strong> protein A with the TNF receptor<br />

TNFR1 (Rahman/McFadden 2006).<br />

Beside other functions, TNF was described to further enhance the expression <strong>of</strong> the<br />

immunoproteasome (Loukissa et al. 2000). In non-pr<strong>of</strong>essional antigen-presenting cells, the<br />

induction <strong>of</strong> immunoproteasome, TAP peptide transporters, and MHC class I complexes <strong>by</strong> TNF<br />

independent <strong>of</strong> IFN-γ activity was observed and in addition increased stability <strong>of</strong> the MHC class I<br />

complexes at the cell surfaces (Hallermalm et al. 2001). Thus, induction <strong>of</strong> TNF in BMM after<br />

IFN-γ treatment might intensify the effect on antigen processing and presentation after a second<br />

stimulus. Also in the serum-free system, the inducibility <strong>of</strong> TNF in IFN-γ and LPS treated BMM was<br />

described (Eske et al. 2009). IL-6, IL-10, and IL-12, which were also inducible <strong>by</strong> IFN-γ and LPS<br />

(Eske et al. 2009), were not expressed in IFN-γ treated BMM (this study), and the IFN-γ and LPS<br />

inducible chemokine Ccl2 (MCP-1; Eske et al. 2009) was expressed but not differentially regulated<br />

in IFN-γ treated BMM (this study). This strongly hints for the explanation that the induction only<br />

occurs when a second stimulus like LPS is given.<br />

After IFN-γ treatment, the induction <strong>of</strong> anti-inflammatory IL-10 receptor (Moore et al. 2001)<br />

as well as the induction <strong>of</strong> Il18bp coding for an IL-18 binding protein, which is a natural<br />

endogenous inhibitor <strong>of</strong> the proinflammatory IL-18 (McInnes et al. 2000, Gracie et al. 2003,<br />

Novick D et al. 1999), was observed (this study). Not only Stat1 and Irf1 <strong>of</strong> the positive IFN-γ<br />

feedback were induced, but also Stat3 and Socs1 which are implicated in negative feedback<br />

(Schroder et al. 2004, Hu et al. 2008). This might indicate the possible reactivity to antiinflammatory<br />

stimuli and a confinement <strong>of</strong> inflammatory response.<br />

After the priming signal <strong>of</strong> IFN-γ, the BMM are not fully activated and are not expected to<br />

secrete cytokines, but rather after a second stimulus (Hu et al. 2008). Analysis <strong>of</strong> cytokine<br />

secretion <strong>of</strong> BMM beyond the examples analyzed until now will be <strong>of</strong> interest, either after<br />

stimulation with LPS, or after stimulation with molecules <strong>of</strong> Gram-positive bacteria or even<br />

infection e. g. with S. aureus, as it will be in focus <strong>of</strong> research for follow-up experiments to this<br />

study.<br />

185


Maren Depke<br />

Discussion and Conclusions<br />

Three other induced genes, kynreninase (Kynu), inducible nitric oxide synthase 2 (Nos2), and<br />

tryptophanyl-tRNA synthetase (Wars), found attention. These are closely linked to inflammation<br />

and interferon action.<br />

Kynureninase was reported to be induced in immortalized murine macrophages (cell line<br />

MT2) <strong>by</strong> IFN-γ, which was interpreted as an activation <strong>of</strong> the kynurenine pathway leading to the<br />

formation <strong>of</strong> quinolinic acid from tryptophan since the cells also induced indoleamine 2,3-<br />

dioxygenase (IDO), which catalyzes the first step <strong>of</strong> tryptophan degradation (Alberati-Giani et al.<br />

1996). Interestingly, BMM after IFN-γ treatment (this study) did not differentially express<br />

indoleamine 2,3-dioxygenase (IDO) which is known to be induced <strong>by</strong> interferons in many cell<br />

types (Carlin et al. 1989, Taylor MW/Feng 1991, Alberati-Giani et al. 1996). Even more,<br />

expression <strong>of</strong> Ido1 and Ido2 was low and even absent in several samples. In literature references,<br />

a lack <strong>of</strong> inducible IDO activity <strong>by</strong> IFN-γ was observed in murine peritoneal macrophages. These<br />

cells were able to kill <strong>pathogen</strong>s, and the antimicrobial effect was not diminished <strong>by</strong> addition <strong>of</strong><br />

tryptophan like it was observed for monocyte-derived macrophages (Murray HW et al. 1989).<br />

It was reported that the immune defense mechanisms <strong>of</strong> IDO/tryptophan degradation and<br />

nitric oxide synthase (NOS)/arginine degradation influence each other. Nitric oxide synthase<br />

catalyzes the first, rate-limiting step <strong>of</strong> arginine degradation to NO and citrulline<br />

(Knowles/Moncada 1994). NO decreases IDO activity in human mononuclear phagocytes <strong>by</strong><br />

interfering with the heme iron located at the active site <strong>of</strong> IDO enzyme (Thomas et al. 1993).<br />

Experiments gave hints for species-specific reactions to IFN-γ since human macrophages induced<br />

IDO and murine macrophages NOS. Nevertheless, in presence <strong>of</strong> inhibitors <strong>of</strong> NO synthesis also<br />

murine macrophages exhibited IDO activity (Thomas et al. 1993). NOS inhibitors in combination<br />

with IFN-γ further increased the induction <strong>of</strong> IDO mRNA in murine macrophages MT2, which was<br />

already induced after IFN-γ treatment alone (Alberati-Giani et al. 1997). Influence <strong>of</strong> NO on IDO<br />

mRNA could not be detected in the human epithelial cell line RT4, but NO led to accelerated<br />

proteosomal degradation <strong>of</strong> IDO protein. The authors stated that the transcriptional regulation <strong>of</strong><br />

IDO might be species or cell type specific and needs to be furthermore analyzed (Hucke et al.<br />

2004). Not only NO and reactive nitrogen species influence IDO, but there are also hints for<br />

influences in the opposite direction. The intermediate 3-hydroxyanthranilic acid from the<br />

kynurenine pathway <strong>of</strong> tryptophan degradation was shown to inhibit the enzymatic activity <strong>of</strong><br />

iNOS and the iNOS mRNA induction <strong>by</strong> IFN-γ and LPS in mouse macrophage RAW 264.7 cells. The<br />

decrease in iNOS mRNA level resulted from inhibition <strong>of</strong> NFκB activation (Sekkaï et al. 1997).<br />

Even when there are still open questions concerning the interplay between IDO and iNOS, and<br />

even when details in their regulation might differ between mouse and human and between the<br />

different cell types, many information support the proposition that in each physiological situation<br />

a cellular decision towards one <strong>of</strong> the two antimicrobial mechanisms has to be taken since each<br />

mechanism compromises the other. Fittingly, during the treatment <strong>of</strong> BMM with IFN-γ the<br />

induction <strong>of</strong> Nos2, inducible nitric oxide synthase 2, was observed in parallel to the lacking<br />

induction/expression <strong>of</strong> IDO (this study).<br />

Wars is known to be regulated <strong>by</strong> IFN-γ (Fleckner et al. 1991, Buwitt et al. 1992, Bange et al.<br />

1992, Rubin et al. 1991). The biological function <strong>of</strong> the induction <strong>of</strong> this metabolic enzyme in<br />

inflammatory situations is discussed in different ways: 1) protection <strong>of</strong> the <strong>host</strong> from self-induced<br />

tryptophan starvation: In an environment <strong>of</strong> tryptophan degradation, the amino acid would be<br />

saved from IDO activity when complexed to tRNA, not accessible for <strong>pathogen</strong>s, but still available<br />

for <strong>host</strong> protein synthesis (Boasso et al. 2005, Murray MF 2003). 2) compensation for increased<br />

186


Maren Depke<br />

Discussion and Conclusions<br />

tryptophanyl-tRNA need: Interferon inducible proteins with tryptophan-enriched sequences<br />

require above average tryptophanyl-tRNA (Xue/Wong 1995). 3) expanded function <strong>of</strong> the<br />

tryptophanyl-tRNA synthetase: Splice variants were shown to act anti-angiogenic (Tolstrup et al.<br />

1995, Otani et al. 2002, Wakasugi et al. 2002).<br />

It has been shown that bone marrow-derived myeloid dendritic cells (Hara et al. 2008) and<br />

astrocytes (Speciale et al. 1989) take up exogenous kynurenine, which would be an explanation<br />

for the induction <strong>of</strong> kynureninase in BMM after treatment with IFN-γ (this study).<br />

It can be speculated that the induction <strong>of</strong> Kynu and Wars together in BMM after treatment<br />

with IFN-γ might indicate a preparation <strong>of</strong> the BMM for their presence in inflamed tissue where<br />

they might encounter a tryptophan depleted environment.<br />

Interferons are described to induce a group <strong>of</strong> GTPases / GTP binding proteins <strong>of</strong> different<br />

families which play a role in antimicrobial defense. Of each <strong>of</strong> the four families described in<br />

literature (MacMicking 2004) examples were found in the BMM data set after IFN-γ treatment, in<br />

part exhibiting the highest induction factors <strong>of</strong> the data set: p47 GTPases (Igtp, Iigp1, Irgm1,<br />

Tgtp), p65 guanylate-binding proteins (Gbp1, Gbp2, Gbp3, Gbp4, Gbp5, Gbp6), Mx proteins (Mx1,<br />

Mx2), and very large inducible GTPases (Gvin1). The observation <strong>of</strong> induction is entirely in<br />

agreement with literature references and impressive in its entity. This group <strong>of</strong> GTPases provides<br />

the <strong>host</strong> with resistance against viral and microbial <strong>pathogen</strong>s <strong>by</strong> cell-autonomous resistance<br />

mechanisms.<br />

GTPase p47 and p65 cellular knockdown studies and experiments using knockout animals led<br />

to a higher susceptibility during infection. These GTPases were described to be membraneassociated<br />

via myristoylation, isoprenylation, or interaction with e. g. Golgi-proteins. GTPases <strong>of</strong><br />

the p47 family were recruited to phagosomes after infection. Thus, they target intracellular,<br />

vacuolarized <strong>pathogen</strong>s <strong>by</strong> a mechanism proposed to remodel the <strong>pathogen</strong>-containing<br />

compartment and to increase the fusion with lysosomes. GTPases <strong>of</strong> the p65 family were<br />

described to be involved in the control <strong>of</strong> virus infections. More recently, a publication reported a<br />

role for p65 GTPases during defense against bacterial and protozoan infections (Shenoy et al.<br />

2007, Taylor GA et al. 2007, Anderson SL et al. 1999, Carter et al. 2005, Degrandi et al. 2007). Mx<br />

proteins, which intracellularly bind to virus particles, are part <strong>of</strong> the innate immune defense<br />

against several RNA viruses. In mice, the location <strong>of</strong> the Mx protein type (nuclear Mx1, cytosolic<br />

Mx2) correlates with the replication site <strong>of</strong> the viruses against which the Mx proteins provide<br />

resistance (Haller/Kochs 2002, Haller et al. 2007). Finally, very large inducible GTPases (VLIG) are<br />

proteins <strong>of</strong> approximately 280 kDa. Despite the big size, the protein is encoded in a single exon.<br />

Mice harbor six different, but similar VLIG genes, which are organized in a gene cluster on<br />

chromosome 7 (Klamp et al. 2003). The genomes <strong>of</strong> primates and carnivores do not include VLIG<br />

sequences, which probably were lost during evolution since other vertebrata own VLIG (Li et al.<br />

2009).<br />

Literature data indicate that C57BL/6 BMM do not synthesize Gbp-1 protein after interferon<br />

induction because they own a different allele <strong>of</strong> the Gbp-1 gene (Staeheli et al. 1984). Although<br />

this study detected a 6-fold increase in Gbp-1 mRNA <strong>of</strong> C57BL/6 BMM after IFN-γ treatment,<br />

even the induced mRNA level was still very low. This change in mRNA was not detectable on<br />

protein level: LC-MS/MS data <strong>of</strong> Dinh Hoang Dang Khoa revealed equal protein abundance in<br />

control and IFN-γ treated C57BL/6 BMM, which was in accordance with the phenotype described<br />

in literature.<br />

187


Maren Depke<br />

Discussion and Conclusions<br />

Only few <strong>genomewide</strong> transcriptome pr<strong>of</strong>iling studies on IFN-γ and macrophages are available<br />

in literature. Kota et al. analyzed the reaction <strong>of</strong> murine RAW 264.7 cells to IFN-γ exposition for<br />

4 h (Kota et al. 2006). Despite the difference <strong>of</strong> macrophage cell line to BMM and <strong>of</strong> treatment<br />

time and IFN-γ dose (4 h and 1000 U/ml vs. 24 h and 300 U/ml) between the published study and<br />

the study described in this thesis, a very good agreement in the differential regulation <strong>of</strong> several<br />

genes was observed (e. g. Oas, GTPases, Socs1, Socs3, Nos2, Cxcl9, Cxcl10, Irf1, Cxcr4,<br />

immunoproteasome and associated genes, MHC molecules, Il18bp, Il10ra, Ctsc, Ptgs2 and<br />

others). Ehrt et al. analyzed BMM after 72 h <strong>of</strong> IFN-γ treatment (100 U/ml), infection for 24 h<br />

with Mycobacterium tuberculosis, or both (Ehrt et al. 2001). The authors observed – in contrast<br />

to the results <strong>of</strong> this study – slightly more repressed than induced genes after IFN-γ treatment<br />

alone. The infection setting resulted in an additional, specific set <strong>of</strong> differentially expressed<br />

genes. Such additional set <strong>of</strong> genes is also expected in case <strong>of</strong> future inclusion <strong>of</strong> further infected<br />

samples in the settings <strong>of</strong> the study described in this thesis. The authors defined a set <strong>of</strong><br />

approximately 1300 genes to be regulated upon treatment <strong>of</strong> BMM with IFN-γ (Ehrt et al. 2001).<br />

This high number is not directly comparable with the results <strong>of</strong> this study, because Ehrt and<br />

colleagues treated each probe set <strong>of</strong> the array as if it represented a single gene. This<br />

simplification was necessary because the analysis was performed before the complete<br />

sequencing <strong>of</strong> the mouse genome was finished. Thus, at that time, the partially overlapping EST<br />

and cDNA sequences were difficult to assign to gene information. Furthermore, the authors<br />

explain the result <strong>of</strong> the high number <strong>of</strong> regulated genes with the long IFN-γ stimulation time <strong>of</strong><br />

72 h. But also between the study <strong>by</strong> Ehrt et al. and the study described in this thesis accordance<br />

was observed (e. g. MHC molecules, Gbp2 and other GTPases, Irg1, Nos2, Cxcl10, Cxcr4).<br />

Interestingly, the authors monitored a strong influence <strong>of</strong> Nos2 deficiency on the gene<br />

expression pr<strong>of</strong>ile after IFN-γ treatment leading to the conclusion that Nos2 directly or indirectly<br />

influences the cellular reaction to IFN-γ stimulus. Zocco and coworkers focused their analyses on<br />

rat hepatic macrophages, i. e. Kupffer cells, stimulated with 1000 U/ml IFN-α or IFN-γ for 8 h<br />

(Zocco et al. 2006). Using an Affymetrix array with about 8800 probe sets, they observed 70<br />

induced and 72 repressed genes <strong>by</strong> IFN-γ, the relation <strong>of</strong> which resembles more that observed <strong>by</strong><br />

Ehrt et al. than the relation determined in the study described in this thesis. Nevertheless, also<br />

here a certain concordance <strong>of</strong> cellular reaction became visible (e. g. Mx, Gbp2, Nos2, Irf1, Stat1,<br />

Oas, immunoproteasome subunits, MHC molecules). In the study <strong>of</strong> Pereira et al. BMM <strong>of</strong> A/J or<br />

BALB/c mice were treated for 18 h with 50 U/ml <strong>of</strong> IFN-γ and analyzed on a 1536 feature cDNA<br />

array from a fetal thymus library (Pereira et al. 2004). For BALB/c BMM, the induction <strong>of</strong> 297<br />

genes and repression <strong>of</strong> 58 genes was recorded. Here, the relation <strong>of</strong> induction to repression is<br />

similar to that observed in the study described in this thesis although the comparison <strong>of</strong> results is<br />

difficult because <strong>of</strong> the very different arrays, which were used in both studies. The comparison<br />

with similar transcriptome studies from literature leads to the conclusion that the results <strong>of</strong> this<br />

study are well supported <strong>by</strong> knowledge on cellular reactions to IFN-γ. In addition, similar to<br />

comparisons <strong>of</strong> other studies, also in this study specific gene expression changes were observed<br />

which might be explained with experimental differences. Nevertheless, for a comprehensive<br />

knowledge <strong>of</strong> IFN-γ effects the study <strong>of</strong> different experimental settings is necessary to which this<br />

study contributes.<br />

During data analysis using the Ingenuity Pathway Analysis tool (IPA, www.ingenuity.com), it<br />

became clear that only a fraction <strong>of</strong> the IFN-γ related, differentially expressed genes were linked<br />

to macrophages or RAW cells in the IPA database. Of about 180 genes, approximately 130 were<br />

188


Maren Depke<br />

Discussion and Conclusions<br />

associated with IFN-γ in other cells, but not in macrophages/RAW cells. Thus, this study provides<br />

a significant contribution to extend the scientific knowledge <strong>of</strong> IFN-γ effects in macrophages.<br />

Strain differences in gene expression occurred between BALB/c BMM and C57BL/6 BMM on<br />

both treatment levels. Most differences were similar at control level and after IFN-γ treatment,<br />

and the differences included genes, which were expressed at a higher level in BALB/c BMM or in<br />

C57BL/6 BMM. Some insights into the data indicate that the differences might be immunerelevant.<br />

For example Cxcl11, the Th1 cell attracting chemokine, differed between the strains<br />

after IFN-γ treatment because it was induced from a common baseline level only in BALB/c BMM.<br />

Since it is known that the immune response <strong>of</strong> C57BL/6 mice is balanced towards Th1 with high<br />

IFN-γ/low IL-4 production in splenocytes after concanavalin A stimulation and that <strong>of</strong> BALB/c<br />

mice towards Th2 with low IFN-γ/high IL-4 production in splenocytes after concanavalin A<br />

stimulation (Mills et al. 2000), the difference after IFN-γ treatment was unexpected. But as the<br />

IFN-γ stimulus was added externally, the difference might display a compensatory variation <strong>of</strong><br />

BALB/c BMM gene expression. C57BL/6 macrophages were reported to produce more NO than<br />

BALB/c macrophages (Mills et al. 2000). The induction <strong>of</strong> Nos2 <strong>by</strong> IFN-γ in C57BL/6 BMM was <strong>by</strong> a<br />

factor <strong>of</strong> about 1.4 stronger than in BALB/c BMM in this study, but the difference was not<br />

significant although a trend for confirmation <strong>of</strong> literature data was visible. Furthermore, BALB/c<br />

macrophages were observed to produce more ornithine/urea because <strong>of</strong> a stronger arginase<br />

activity (Mills et al. 2000). In this study, BALB/c BMM exhibited higher expression <strong>of</strong> arginase 2<br />

than C57BL/6 at control and at IFN-γ treated level as indicated <strong>by</strong> literature data.<br />

The phenotypical differences between the reaction <strong>of</strong> BALB/c and C57BL/6 BMM were <strong>of</strong>ten<br />

determined in the presence <strong>of</strong> IFN-γ and a second stimulus like LPS or infection (Breitbach et al.<br />

2006, Eske et al. 2009). To elucidate molecular reasons for the observed differences in killing <strong>of</strong><br />

<strong>pathogen</strong>s or cytokine production, the inclusion <strong>of</strong> samples subjected to a second stimulus in<br />

addition to IFN-γ is recommended.<br />

The experimental setting <strong>of</strong> this study was planned with the aim to analyze basic principles <strong>of</strong><br />

murine BMM: the reaction to the priming signal IFN-γ on a molecular level and differences <strong>of</strong><br />

reaction between the BMM <strong>of</strong> both mouse strains. To further elucidate reasons for the different<br />

success <strong>of</strong> the mouse strains to fight and overcome infections in vivo and in vitro, experiments<br />

will need to be expanded to the analysis <strong>of</strong> IFN-γ treatment in combination with bacterial<br />

infection. This might include studies using Burkholderia infections for which attenuated mutants<br />

exist <strong>of</strong> whose study interesting results are anticipated. Furthermore, the first proteome analyses<br />

were performed for infection experiments <strong>of</strong> BMM with S. aureus, <strong>of</strong> which the first results are<br />

presented in the thesis <strong>of</strong> Dinh Hoang Dang Khoa. Further complementing results are expected<br />

from the extension <strong>of</strong> the analysis to transcriptome level.<br />

189


Maren Depke<br />

Discussion and Conclusions<br />

HOST CELL GENE EXPRESSION PATTERN IN AN<br />

IN VITRO INFECTION MODEL<br />

Several different cell types <strong>of</strong> mammalian <strong>host</strong>s are in the risk <strong>of</strong> encounter with <strong>pathogen</strong>s<br />

depending <strong>of</strong> the entry and manifestation site <strong>of</strong> infection. These are immune cells as well as cells<br />

associated with structural and functional aspects <strong>of</strong> the tissue. In this study, the human bronchial<br />

epithelial cell line S9 (American Type Culture Collection ATCC, Manassas, VA, USA; www.atcc.org;<br />

S9 cell ATCC number CRL-2778) served as a model system to study the reaction <strong>of</strong> epithelial<br />

cells to infection. With regard to the identification <strong>of</strong> S. aureus as one <strong>of</strong> the leading causative<br />

organisms <strong>of</strong> pneumonia (Goto et al. 2009), this species was employed as model <strong>pathogen</strong>. More<br />

specific, S. aureus RN1HG, a rsbU + repaired RN1-derivative strain (Herbert et al. 2010) with a<br />

SigB-positive phenotype, was chosen.<br />

The infection setting involved bacterial cultivation in an adapted cell culture medium (Schmidt<br />

et al. 2010), which allowed inoculation <strong>of</strong> eukaryotic <strong>host</strong> cell cultures with a fraction <strong>of</strong> complete<br />

bacterial culture. This experimental setup allowed the study <strong>of</strong> <strong>host</strong> cell reactions to the influence<br />

<strong>of</strong> both bacterial factors, its secreted proteins from supernatant and the membrane-bound<br />

factors, which both interact during infection and contribute to the success <strong>of</strong> the bacterial cells. It<br />

also avoided potentially disturbing influences <strong>of</strong> bacterial cell handling like that occurring during<br />

centrifugation and washing.<br />

In a combined approach <strong>of</strong> transcriptome (Maren Depke) and proteome (Melanie Gutjahr)<br />

analysis, the <strong>host</strong> reaction to infection and bacterial internalization was recorded. Two time<br />

points, 2.5 h and 6.5 h after start <strong>of</strong> infection, were selected for sampling. Because only about<br />

50 % <strong>of</strong> <strong>host</strong> cells in infected cell culture plates harbor internalized staphylococci after infection,<br />

the fraction <strong>of</strong> non-infected cells was removed <strong>by</strong> FACS-sorting, which was feasible because a<br />

GFP-expressing S. aureus RN1HG strain has been used. The remaining infected S9 cells were<br />

compared to medium control cells.<br />

In the comparison <strong>of</strong> infected S9 cell proteome and transcriptome signatures, a considerable<br />

time shift <strong>of</strong> cellular reaction between both analysis levels was evident. In samples <strong>of</strong> the first<br />

time point 2.5 h, approximately half the number <strong>of</strong> proteins differed in abundance between<br />

infected and control cells in comparison to the number <strong>of</strong> regulated proteins at the later time<br />

point <strong>of</strong> 6.5 h. Conversely, for the transcriptome analysis, the number <strong>of</strong> differentially expressed<br />

genes at the first time point amounted to about 3.5 % <strong>of</strong> the number <strong>of</strong> regulated genes at the<br />

second time point. Thus, the reaction on proteome level started to a stronger extent between<br />

inoculation and first analysis time point 2.5 h, whereas the transcriptional reaction most<br />

articulately started between the first (2.5 h) and the second (6.5 h) time point.<br />

The cellular reaction to infection in this experimental setting seems to lead first to protein<br />

abundance changes independent from mRNA changes, which on the one hand probably rely on<br />

the already existing messengers and on the other hand might represent post-transcriptional,<br />

translational and degradation/turnover processes as reduction in protein abundance prevailed at<br />

the early time point. Differentially expressed genes and regulated proteins did not overlap at the<br />

first time point 2.5 h after start <strong>of</strong> infection. Two genes, IFIT2 and IFIT3, were induced at both the<br />

190


Maren Depke<br />

Discussion and Conclusions<br />

2.5 h and the 6.5 h time points, but the protein abundance increased only at the 6.5 h time point.<br />

This indicates that the first mRNA expression changes were manifested on proteome level<br />

probably after the first analyzed time point (2.5 h) and before or at latest directly at the second<br />

analyzed time point (6.5 h).<br />

A time shift between detectable messenger and protein level changes can explain the<br />

observation that approximately 5 % <strong>of</strong> protein abundance changes found their correspondence at<br />

transcript level at the 6.5 h time point, and vice versa, the even smaller fraction <strong>of</strong> 0.8 % <strong>of</strong><br />

differential expression was realized on proteome level at the same time point. It has to be<br />

mentioned that a small number <strong>of</strong> genes/proteins showed reverse effects in transcriptome and<br />

proteome analysis, or changes in protein abundance preceeded mRNA changes. This can be<br />

explained <strong>by</strong> post-transcriptional, translational and protein turnover mechanisms, <strong>by</strong> time shift<br />

aspects or <strong>by</strong> the possibility <strong>of</strong> counter-regulatory processes. During proteome analysis, different<br />

methods have to be applied to study proteins from different compartments or with different<br />

physical characteristics. In this study, a gel-free proteomics approach was applied, which is<br />

known to yield less information on membrane proteins because these are not accessed during<br />

extract preparation and enzymatic digestion. Thus, a defined part <strong>of</strong> the microarray results can<br />

practically not match proteome results for lack <strong>of</strong> accessibility. The results from the comparison<br />

<strong>of</strong> the proteome and the transcriptome analysis approach underline the importance <strong>of</strong><br />

performing both analysis levels in parallel. Information <strong>of</strong> both approaches complement each<br />

other and supply valuable data to establish a comprehensive view <strong>of</strong> the experimental system.<br />

Although transcriptome pr<strong>of</strong>iling revealed the very small number <strong>of</strong> 40 differentially<br />

expressed genes in S9 cells 2.5 h after start <strong>of</strong> infection with S. aureus RN1HG, these gene<br />

expression changes indicated the beginning response <strong>of</strong> the <strong>host</strong> cells to infection. Furthermore,<br />

these changes were in very good agreement with the differential gene expression at the second<br />

analysis time point four hours later.<br />

At the 2.5 h time point, a strong induction <strong>of</strong> a proinflammatory response was observed. This<br />

included cytokines like IL-6 and IFN-β, genes related to inflammation mediators like PTGS2, and<br />

genes which are linked to these aspects because they are part <strong>of</strong> signal transduction and<br />

transcription processes. In addition to the observation <strong>of</strong> proinflammatory mechanisms, gene<br />

expression changes which aim to counter-regulate and balance the inflammation were observed<br />

already 2.5 h after start <strong>of</strong> infection. Nevertheless, the infection influenced the <strong>host</strong> cells very<br />

strongly, and hints for morphological changes and a limitation <strong>of</strong> cell cycle progression became<br />

visible.<br />

The study included a second sampling point 6.5 h after start <strong>of</strong> infection. The cellular state at<br />

this time point was characterized <strong>by</strong> a stronger loss <strong>of</strong> viability, but these non-viable cells were<br />

not included in the sample preparation after FACS-sorting. Thus, a viable, GFP-positive infected<br />

cell fraction was utilized for transcriptome pr<strong>of</strong>iling. Resulting differentially expressed genes<br />

distinguished a rather strong cellular response as the number <strong>of</strong> regulated genes increased <strong>by</strong> a<br />

factor <strong>of</strong> approximately 30 in comparison to the first time point.<br />

In a general view, the S9 cells started a regulatory gene expression program 6.5 h after start <strong>of</strong><br />

infection, which was strongly associated with diverse inflammation and cell death related<br />

functions. These included signal transduction linked genes for the interferon signaling cascade<br />

together with IFN-β itself, which has already been induced at the 2.5 h time point, pattern<br />

recognition receptors, antigen presentation and immunoproteasome, but also death receptor<br />

signaling. The amplification <strong>of</strong> cellular response was indicated <strong>by</strong> the emergence <strong>of</strong> functional<br />

191


Maren Depke<br />

Discussion and Conclusions<br />

associated groups <strong>of</strong> genes, <strong>of</strong> which few were regulated at the early and several at the later time<br />

point <strong>of</strong> analysis. Examples are given with leukemia inhibitory factor LIF (induced at both time<br />

points) and its receptor (induced after 6.5 h), ligand CD274 (PD-L1, induced at both time points)<br />

and PD-L2 (induced after 6.5 h), several interferon regulated genes at both time points or<br />

additionally at the 6.5 h time point, and the histone genes, <strong>of</strong> which one was repressed at the<br />

2.5 h time point and further 17 were added at the 6.5 h time point.<br />

Furthermore, indoleamine 2,3-dioxygenase 1 (IDO1) belonged to the genes with the highest<br />

induction factors at the 6.5 h time point. The gene codes for a key molecule in<br />

immunomodulation, microbial growth control, and <strong>pathogen</strong> immune escape (Puccetti 2007,<br />

Zelante et al. 2009). IDO is known to be induced <strong>by</strong> interferons (Carlin et al. 1989,<br />

Taylor MW/Feng 1991) e. g. during infection and inflammation, but also in conditions like stress,<br />

which result in a transient increase <strong>of</strong> proinflammatory cytokines (Kiank et al. 2010). The enzyme<br />

catalyzes the reaction <strong>of</strong> tryptophan to formylkynurenine. This is the first rate-limiting step <strong>of</strong> the<br />

so-called “kynurenine pathway”, which leads to NAD + biosynthesis or to the complete oxidative<br />

degradation <strong>of</strong> kynurenine (King/Thomas 2007, Xie et al. 2002). One <strong>of</strong> the main effects <strong>of</strong><br />

induced IDO activity is the depletion <strong>of</strong> tryptophan, which is discussed to be a mechanism to<br />

inhibit microbial growth (Pfefferkorn 1984, Byrne et al. 1986, Müller et al. 2009). But this<br />

mechanism can be counteracted <strong>by</strong> bacteria, which might induce their tryptophan biosynthesis<br />

or which developed own counter-regulatory mechanisms. An example is Chlamydophila psittaci<br />

which owns adapted tryptophan biosynthesis genes that enable the bacterium to produce<br />

tryptophan from kynurenine taken from their <strong>host</strong>’s metabolism (Xie et al. 2002).<br />

During metabolism <strong>of</strong> tryptophan, several neuro- and immune-active substances are<br />

produced, <strong>of</strong> which some additionally exhibit toxic features and thus might act antimicrobially.<br />

The intermediates <strong>of</strong> tryptophan and kynurenine degradation, 3-hydroxykynurenine and α-<br />

picolinic acid, were reported to inhibit growth <strong>of</strong> S. aureus and other bacterial species in a murine<br />

transplantation model even in the presence <strong>of</strong> tryptophan (Saito et al. 2008, Narui et al. 2009).<br />

The most prominent characteristic <strong>of</strong> IDO is the induction <strong>of</strong> an anti-inflammatory bias after a<br />

proinflammatory stimulus. This effect aims to limit inflammation and the accompanying damage<br />

to the <strong>host</strong> tissue. It has to be considered that the experimental setting in this study only<br />

included one single cell type, the bronchial epithelial cell line S9. Thus, interpretation <strong>of</strong> an antiinflammatory<br />

effect assumes that bronchial epithelial cells might also in vivo induce IDO, e. g.<br />

during staphylococcal pneumonia. Tryptophan degradation intermediates were shown to inhibit<br />

proliferation <strong>of</strong> CD4 + and CD8 + T cells and <strong>of</strong> NK cells (Frumento et al. 2002). Other experiments<br />

supported a proposed mechanism <strong>of</strong> inhibition <strong>of</strong> T cell proliferation, which was triggered <strong>by</strong> the<br />

low concentration <strong>of</strong> tryptophan resulting from IDO activity and mediated <strong>by</strong> GCN2 kinase<br />

(EIF2AK4, eukaryotic translation initiation factor 2 alpha kinase 4), a sensor for uncharged tRNA<br />

(Munn et al. 2005). The contradiction that tryptophan depletion leads to an antimicrobial effect<br />

on the one hand and to an inhibition <strong>of</strong> T cell mediated immune response on the other hand was<br />

resolved <strong>by</strong> Müller and coworkers, who determined the concentration limit necessary for the two<br />

effects. They observed that bacterial inhibition (50 % inhibitory concentration <strong>of</strong> 1.9 µM) took<br />

already place at a higher tryptophan concentration whereas tryptophan had to be depleted even<br />

further (50 % inhibitory concentration <strong>of</strong> 0.1 µM) to achieve T cell inhibition (Müller et al. 2009).<br />

Besides the described aspects, in an in vivo infection situation dendritic cells (DC) will contribute<br />

to the immune response. Further IDO-mediated immune-modulation mechanisms involve DC<br />

mediated T cell tolerance (Popov/Schultze 2008). The induction <strong>of</strong> IDO in epithelial cells near<br />

192


Maren Depke<br />

Discussion and Conclusions<br />

sites <strong>of</strong> infection might help to constrain the immune defense reactions to the actually infected<br />

cells and to protect the non-infected neighboring cells from damage <strong>by</strong> immune cells like<br />

cytotoxic T cells (King/Thomas 2007). This confinement <strong>of</strong> immune response evolutionarily<br />

developed in favor <strong>of</strong> the <strong>host</strong>. Nevertheless, in certain context it benefits <strong>pathogen</strong>s, which<br />

might cause persistent infections since the immune system switched to an anti-inflammatory<br />

state (Zelante et al. 2009).<br />

In relation to IDO1, the increased expression <strong>of</strong> tryptophanyl-tRNA synthetase (WARS) was<br />

conspicuous, which was also confirmed on protein level. In contrast to bacterial tRNA<br />

synthetases, the eukaryotic or mammalian enzymes were studied to a lesser extent. The enzyme<br />

was identified and described to be IFN-γ induced <strong>by</strong> two groups, Fleckner and coworkers as well<br />

as Buwitt, Bange and colleagues almost 20 years ago (Fleckner et al. 1991, Buwitt et al. 1992,<br />

Bange et al. 1992) and separately, the inducibility <strong>by</strong> IFN-α and IFN-γ was demonstrated (Rubin et<br />

al. 1991). The protein domain structure (reviewed <strong>by</strong> Kisselev 1993) as well as the gene’s exonintron-organization<br />

and interferon response elements are long-known (Frolova et al. 1993). It<br />

was hypothesized that the induction <strong>of</strong> tryptophanyl-tRNA synthetase in parallel to the<br />

tryptophan catabolizing enzyme IDO aims to protect the <strong>host</strong> from the tryptophan starvation<br />

which is induced in response to proinflammatory stimuli. Tryptophan would be safe from<br />

degradation when complexed to tRNA and still available for <strong>host</strong> protein synthesis (Boasso et al.<br />

2005, Murray MF 2003). Furthermore, the induction <strong>of</strong> tryptophanyl-tRNA synthetase <strong>by</strong><br />

interferon was described to find its biological rationale in the enrichment <strong>of</strong> tryptophan in<br />

immune response related and equally interferon inducible proteins like MHC molecules<br />

(especially in the Ig-domains). Hence, induction <strong>of</strong> tryptophanyl-tRNA synthetase was suggested<br />

to reflect the heightened need <strong>of</strong> tryptophanyl-tRNA to sustain the ability to synthesize these<br />

proteins (Xue/Wong 1995). More recent literature data assign tRNA synthetases expanded<br />

functions (Yang XL et al. 2004). Proteolytic tryptophanyl-tRNA synthetase fragments (Favorova et<br />

al. 1989), splice variants (Tolstrup et al. 1995), and transcription starting from a newly identified<br />

IFN-γ sensitive promoter (Liu J et al. 2004) were reported. Smaller tryptophanyl-tRNA synthetase<br />

variants were shown to act anti-angiogenic (Otani et al. 2002, Wakasugi et al. 2002).<br />

In infected S9 cells 6.5 h after start <strong>of</strong> infection, expression changes <strong>of</strong> cytokine genes,<br />

GTPase/GTP binding proteins, genes involved in (anti)coagulation, anti-oxidant defense, and<br />

complement system, lysosomal proteins and adhesins were obvious. This time point also<br />

included repression <strong>of</strong> different branches <strong>of</strong> de novo lipogenesis like cholesterol, unsaturated<br />

fatty acid, and storage lipid biosynthesis. Fatty acid synthase FASN was repressed in<br />

transcriptome analysis and was equally reduced in abundance in proteome analysis. Contrarily,<br />

several genes which are involved in lipid messenger generation were induced. The hints for<br />

induced ceramide/sphingosine biosynthesis link the induction <strong>of</strong> lipid messenger generating<br />

genes to the observation <strong>of</strong> induced expression <strong>of</strong> genes associated with death receptor signaling<br />

and apoptosis.<br />

The transcriptome pr<strong>of</strong>iling <strong>of</strong> infected S9 cells revealed that in the death receptor signaling<br />

cascade, the receptor FAS and some initiating caspases were induced in parallel with the ligands<br />

TNFSF10 and TNFSF15 and signal transduction genes. But also a caspase inhibitor (CFLAR) and an<br />

anti-apoptotic gene (BAG1) were induced, and CASP9, link to the mitochondrial apoptosis<br />

pathway, was repressed. Among further pro-apoptotic genes, a set <strong>of</strong> five apolipoprotein L genes<br />

(APOL1, APOL2, APOL3, APOL4, APOL6) was detected with induced expression, <strong>of</strong> which one,<br />

APOL2, was also increased in protein abundance. The APOL genes were <strong>of</strong> special interest. The<br />

193


Maren Depke<br />

Discussion and Conclusions<br />

induced APOL1 belongs to the group <strong>of</strong> BH3-only proteins which achieve their apoptotic effect <strong>by</strong><br />

binding to proteins <strong>of</strong> the Bcl-2 family. It is discussed that the apoptotic effect is mediated <strong>by</strong> a<br />

mechanism <strong>of</strong> activating pro-apoptotic members <strong>of</strong> this family or inactivating pro-survival<br />

members in a de-repression mode (Bouillet/Strasser 2002, Adams 2003, Fletcher/Huang 2006).<br />

Accumulation <strong>of</strong> apolipoprotein L1 leads to autophagy (Wan et al. 2008, Zhaorigetu et al. 2008)<br />

and is postulated to be linked to apoptosis (Vanhollebeke/Pays 2006). APOL genes were reported<br />

to be induced in proinflammatory environment (Smith/Malik 2009). Unlike the secreted APOL1,<br />

the other four induced apolipoprotein L genes APOL2, APOL3, APOL4, and APOL6 do not possess<br />

a secretion signal peptide and therefore are thought to be localized intracellularly (Duchateau et<br />

al. 1997, Page et al. 2001). Like APOL1, they also contain BH3-domains and are supposed to be<br />

associated with programmed cell death and immune response (Liu Z et al. 2005).<br />

Most interestingly, a second function <strong>of</strong> APOL1 as human serum factor which lyses serumsensitive<br />

Trypanosoma species was identified (Vanhamme et al. 2003). After uptake, APOL1<br />

forms pores in the trypanosomal lysosomes, and the subsequent lysosomal disruption leads to<br />

killing <strong>of</strong> the parasite (Pérez-Morga et al. 2005). The species Trypanosoma brucei rhodesiense is<br />

serum-resistant. Its antagonistic protein SRA targets APOL1, inhibits the lytic function, and thus<br />

helps the <strong>pathogen</strong> to evade the immune response (Xong et al. 1998, De Greef et al. 1989,<br />

Lecordier et al. 2009).<br />

The APOL1 domain which is responsible for APOL1-SRA binding is called SID, and the other<br />

APOL proteins exhibit homologous regions to this domain. The region was subjected to rapid<br />

evolution in all six APOL proteins, and another region – called MAD – evolved rapidly only in<br />

APOL6. The authors Smith and Malik predict from these results the existence <strong>of</strong> further unknown<br />

antagonists to these regions, especially in intracellular <strong>pathogen</strong>s, which might take advantage<br />

from inhibition <strong>of</strong> <strong>host</strong> cell death (Smith/Malik 2009).<br />

At the 6.5 h post-infection time point, S9 cells induced the expression <strong>of</strong> several chemokines<br />

and cytokines. This set included CCL2 and CCL5, CXCL10, IFNB1, IL6, IL12A and others, and was<br />

interpreted as a pro-inflammatory response. Literature references also report the induction <strong>of</strong><br />

chemokines/cytokines <strong>by</strong> epithelial cells after a challenge with S. aureus with the aim to recruit<br />

immune cells and activate innate and adaptive immune responses (Moreilhon et al. 2005,<br />

Peterson ML et al. 2005). Also endothelial cells exhibit a similar response to infection with<br />

S. aureus (Matussek et al. 2005). The induced chemokines/cytokines from this study and the<br />

published references overlap partly (e. g. IL6, IL15), but still contain specifically regulated genes.<br />

The experiments were performed with different S. aureus strains. They probably influence the<br />

<strong>host</strong> gene expression to a different extent depending on their differing repertoire <strong>of</strong> virulence<br />

factors as it was shown for different S. aureus strains infecting endothelial cells (Grundmeier et<br />

al. 2010), and – more distantly related – also in plasma cytokine pr<strong>of</strong>iles <strong>of</strong> patients suffering<br />

from Gram-positive or Gram-negative sepsis and in microarray data sets <strong>of</strong> LPS or heat-killed<br />

S. aureus Cowan ex vivo stimulated whole blood samples (Freezor et al. 2003). Possibly also the<br />

different <strong>host</strong> cell lines have different specificities for the induction <strong>of</strong> pro-inflammatory<br />

cytokines. Despite the difference, in conformity between the different studies the epithelial cells<br />

were recognition sites for infection and mediators <strong>of</strong> this information to the immune system.<br />

Since the effect was very explicit they were even termed to be components <strong>of</strong> the innate immune<br />

system (Peterson ML et al. 2005).<br />

Infected S9 cells effectuate a clearly immune defense associated transcriptomic response to<br />

infection. Interestingly, more than 100 genes <strong>of</strong> the infection-regulated set in S9 cells were also<br />

194


Maren Depke<br />

Discussion and Conclusions<br />

found in the IFN-γ dependently regulated genes in BALB/c or C57BL/6 macrophages, when the<br />

gene list was translated to murine homologues in Rosetta Resolver s<strong>of</strong>tware with the help <strong>of</strong> the<br />

EntrezGene HomoloGene database (http://www.ncbi.nlm.nih.gov/homologene). Some genes<br />

even fit together in their regulation in both studies: S9 cells induced cytokine IL12A and<br />

indoleamine 2,3-dioxygenase 1 (IDO1) upon infection, BMM induced the receptor Il12rb1 and<br />

kynureninase (Kynu) after activation <strong>by</strong> IFN-γ.<br />

An important mediator <strong>of</strong> inflammation is prostaglandin E 2 (PGE 2 ) whose production is mainly<br />

mediated <strong>by</strong> prostaglandin-endoperoxide synthase (PTGS), which generates the inflammation<br />

mediators prostaglandin (PG) G 2 and H 2 from arachidonic acid. From PGH 2 , the other<br />

prostaglandins like PGE 2 are formed <strong>by</strong> different synthases (Spencer et al. 1998, Steer/Corbett<br />

2003, Park JY et al. 2006). In this study, PTGS2 and PGE 2 receptors PTGER4 and PTGER2 were<br />

induced at least at one analyzed time point. The mechanism <strong>of</strong> PTGS2 induction involves among<br />

others NFκB (Newton et al. 1997, Lin et al. 2000, Tsatsanis et al. 2006). More specifically, it was<br />

published that lipoteichoic acid (LTA) from S. aureus induced PTGS2 protein in human pulmonary<br />

epithelial A549 cells and that this also led to PGE 2 production to which phospholipase A 2 (PLA2)<br />

contributed (Lin et al. 2001). The authors demonstrate an induction <strong>of</strong> PTGS2 via a mechanism in<br />

which LTA first activates phosphatidylcholine-phospholipase C or D (PC-PLC, PC-PLD), whose<br />

product diacylglycerol (DAG) subsequently activates protein kinase C (PKC), which finally leads to<br />

the activation <strong>of</strong> NFκB and NFκB-dependent PTGS2 (Lin et al. 2001).<br />

In S. aureus RN1HG infected S9 cells, the induction <strong>of</strong> phospholipase A 2 (PLA2G4C),<br />

phospholipase C (PLCB4, PLCG2, PLCH1), protein kinase C µ (PRKD1, PRKD2), and PTGS2 was<br />

observed, which fits very well to the literature data. As it is known that NFκB activation results in<br />

autoregulation and induction <strong>of</strong> its inhibitors (Sun et al. 1993, Le Bail et al. 1993, Liptay et al.<br />

1994, Eto et al. 2003, Trinh et al. 2008), the induction <strong>of</strong> NFKBIZ can be regarded as indirect hint<br />

for NFκB activity in infected S9 cells.<br />

Airway epithelial cells have been used to study S. aureus in vitro infection before. In a farreaching<br />

microarray and RT-PCR study, Moreilhon and coworkers analyzed the reaction <strong>of</strong> the<br />

human airway glandular cell line MM-39 to the S. aureus 8325-4 strain in two settings: First,<br />

diluted bacterial supernatants were added to the <strong>host</strong> cell culture. In a second experiment, PBSwashed<br />

bacterial cells were used to infect the eukaryotic cell culture (Moreilhon et al. 2005).<br />

Bacteria were cultivated in TSB, and when a concentration <strong>of</strong> 5E+08 cfu/ml was reached, cells or<br />

supernatants were used for the infection experiments. This concentration corresponds to a time<br />

point during exponential growth. Inoculation was performed with 10 % supernatant for a time<br />

span from 1 h to 24 h or with a MOI <strong>of</strong> 50 for infection with viable staphylococci for 3 h. Thus,<br />

additional to a different <strong>host</strong> cell line and S. aureus strain – whose known SigB-negative<br />

phenotype (Kullik/Giachino 1997) probably is the main difference to the phenotypically SigBpositive<br />

strain RN1HG – the experimental setting differed from the one used in this study. This<br />

necessarily has impact on the comparability <strong>of</strong> the results. Furthermore, the strain 8325-4 was<br />

observed to produce a lipase and protease sensitive lipoprotein inhibitor <strong>of</strong> internalization into<br />

endothelial cells. Reduced internalization consequently led to a reduction <strong>of</strong> the endothelial cells’<br />

cytokine production (Yao et al. 2000).<br />

Moreilhon et al. observed distinct gene expression pr<strong>of</strong>iles <strong>of</strong> supernatant and viable<br />

staphylococci treated <strong>host</strong> cells in which bacterial supernatant led to stronger alterations (higher<br />

number as well as stronger magnitude) than washed viable staphylococci.<br />

195


Maren Depke<br />

Discussion and Conclusions<br />

The comparison between the published bacterial supernatant effects and the infection effects<br />

recorded in this study revealed similar functional aspects, although these functions were not<br />

always represented <strong>by</strong> the same examples. Nevertheless, both studies confirmed each other: 1)<br />

NFκB action (e. g. IL6, JUNB, NR4A2), 2) inflammation (e. g. LIF), 3) inflammatory mediator-related<br />

genes (e. g. PTGS2 alias COX-2). Inflammation-related signal transduction processes were present<br />

with different members <strong>of</strong> the same families. Moreilhon et al. found induction <strong>of</strong> JAK1, STAT1,<br />

STAT3 and SOCS2, whereas this study detected induction <strong>of</strong> JAK2, STAT2 and SOCS1. Similarly,<br />

Moreilhon et al. described induction <strong>of</strong> pro-inflammatory chemokines/interleukins CXCL1, CXCL2,<br />

CXCL3, CCL20, IL-1α, IL-1β, IL-8, IL-20 and IL-24, and this study found a different set, namely CCL2,<br />

CCL5, CSF1, CXCL10, CXCL11, CXCL16, IL-7, IL-12, IL-15, IL-28, IL-29, and in addition IFNB. From the<br />

group <strong>of</strong> toll-like receptors, TLR2 was induced in the experiments <strong>of</strong> Moreilhon et al. while TLR3<br />

was detected with increased expression in the study described in this thesis.<br />

Also cell cycle/apoptosis related genes were found in both studies, but with different<br />

examples. Moreilhon et al. could determine a necrotic phenotype with a transient apoptosis<br />

phase 8 h to 10 h after <strong>host</strong>-<strong>pathogen</strong> interaction. In this study, the final fate <strong>of</strong> the infected cells<br />

was difficult to judge on, because the functional/physiological measurements have not been<br />

performed yet. Here, only an arrest <strong>of</strong> growth could be postulated from the transcriptome data.<br />

Nevertheless, this is in agreement with proteome result interpretation <strong>of</strong> Melanie Gutjahr.<br />

Also another study using the cell line MM-39 and washed S. aureus 8325-4 cells from<br />

stationary growth phase (da Silva et al. 2004) describes the epithelial cell reaction as necrotic cell<br />

death after a phase <strong>of</strong> apoptosis, where a higher concentration <strong>of</strong> bacterial cells accelerated the<br />

begin <strong>of</strong> necrosis. Finally, after 24 h a very strong damage <strong>of</strong> <strong>host</strong> cells was observed.<br />

Nevertheless, the authors observed that MM-39 <strong>host</strong> cells were able to defend themselves in the<br />

early phase <strong>of</strong> infection <strong>by</strong> the production <strong>of</strong> SLPI, secretory leukocyte peptidase inhibitor, which<br />

was exhausted in later phases <strong>of</strong> infection. It has to be mentioned that extracellular bacteria<br />

were not killed in the experiments <strong>of</strong> da Silva and coworkers and that the strong replication <strong>of</strong><br />

non-internalized and <strong>host</strong>-cell escaped bacteria in the medium during the infection assay and the<br />

parallel production <strong>of</strong> toxic bacterial products might have influenced the <strong>host</strong> cells’ reactions<br />

(da Silva et al. 2004). However, transcriptome data <strong>of</strong> S9 cells during the analysis window <strong>of</strong> the<br />

study described in this thesis do not allow confirmation <strong>of</strong> an active defense strategy using SLPI:<br />

SLPI mRNA was expressed, but at a low level, and differential expression was not observed after<br />

infection. Possibly the bronchial epithelial cell line S9 has a different gene expression repertoire<br />

from that <strong>of</strong> the airway glandular cell line MM-39.<br />

In summary, the in vitro infection study <strong>of</strong> human bronchial epithelial S9 cells and S. aureus<br />

RN1HG resulted in a picture <strong>of</strong> a fast, strong proinflammatory reaction <strong>of</strong> <strong>host</strong> cells which<br />

revealed central immune response related processes. In agreement with literature reports, the<br />

data led to the conclusion that epithelial cells act as recognition sites for infection and pass this<br />

information to immune cells. Nevertheless, the comparison with other studies revealed that<br />

especially each <strong>host</strong> cell – bacterial strain combination, but also the additional experimental<br />

settings provoke a specific aspect <strong>of</strong> <strong>host</strong> response which supports and necessitates the<br />

extension <strong>of</strong> research to further <strong>of</strong> these combinations. Furthermore, the comparison <strong>of</strong><br />

proteome and transcriptome results emphasized that the application <strong>of</strong> both complementing<br />

approaches is recommended and strongly helps to achieve a comprehensive view <strong>of</strong> <strong>host</strong><strong>pathogen</strong><br />

<strong>interactions</strong> on molecular level.<br />

196


Maren Depke<br />

Discussion and Conclusions<br />

PATHOGEN GENE EXPRESSION PROFILING<br />

Beside its characteristics as commensal colonizer as well as origin <strong>of</strong> a variety <strong>of</strong> infectious<br />

diseases, S. aureus was identified as one <strong>of</strong> the leading causative organisms <strong>of</strong> pneumonia<br />

besides Streptococcus pneumonia and Haemophilus influenzae (Goto et al. 2009). The <strong>pathogen</strong><br />

is easily transferred to the lung, e. g. <strong>by</strong> aspiration or medical devices, where it encounters<br />

immune cells as well as cells associated with structural and functional aspects <strong>of</strong> the lung like<br />

epithelial cells.<br />

In this study, the human bronchial epithelial cell line S9 was applied to study the <strong>interactions</strong><br />

<strong>of</strong> epithelial cells with S. aureus RN1HG. Here, the analysis <strong>of</strong> the <strong>pathogen</strong> expression pr<strong>of</strong>ile<br />

complements the analysis <strong>of</strong> <strong>host</strong> cell expression patterns, which has been described before.<br />

Similar as in the other studies described in this thesis, also the internalized staphylococci were<br />

monitored in a combined approach <strong>of</strong> transcriptome (Maren Depke) and proteome (Sandra<br />

Scharf) analysis. Internalized staphylococci were extracted from their S9 <strong>host</strong> cells and the<br />

bacterial RNA pr<strong>of</strong>ile was recorded using a tiling array approach. Bacterial intracellular proteins<br />

were monitored and quantified after stable isotope labeling with amino acids in cell culture<br />

(SILAC), FACS-sorting and mass spectrometric analysis.<br />

The advantage <strong>of</strong> the experimental settings in this study was the application <strong>of</strong> complete<br />

bacterial culture with both secreted proteins from supernatant and whole bacterial cells to the<br />

<strong>host</strong> S9 cell culture. Such experimental design was improved after establishment <strong>of</strong> an adapted<br />

cell culture medium named pMEM (Schmidt et al. 2010) which allows reproducible bacterial<br />

growth, but avoids the inoculation <strong>of</strong> <strong>host</strong> cell culture with highly artificial established bacterial<br />

culture media like TBS. Hence, the study <strong>of</strong> <strong>host</strong>-<strong>pathogen</strong> <strong>interactions</strong> in the context <strong>of</strong> all<br />

bacterial factors, membrane-bound and secreted, and additionally without any influence on<br />

bacterial physiology <strong>by</strong> prolonged handling, centrifugation and washing <strong>of</strong> bacteria was<br />

performed using exponential growth phase cultures <strong>of</strong> the rsbU + repaired RN1-derivative strain<br />

S. aureus RN1HG (Herbert et al. 2010).<br />

Additional information on gene expression pattern <strong>of</strong> S. aureus RN1HG during stationary<br />

growth phase in the pMEM medium was available from a second study, which dealt with the<br />

comparison <strong>of</strong> gene expression pattern in different growth media. That study was part <strong>of</strong> an<br />

international co-operation in the settings <strong>of</strong> the EU-IP-FP6-project BaSysBio (LSHG-CT2006-<br />

037469) consortium. Here, proteome analysis was not included.<br />

Both gene expression studies, <strong>of</strong> medium comparison and <strong>of</strong> the infection experiment, applied<br />

custom tiling arrays produced and processed <strong>by</strong> NimbleGen.<br />

First, the knowledge on the gene expression pattern <strong>of</strong> stationary growth phase was used to<br />

determine the most suitable control sample group for the infection experiment study.<br />

In stationary growth phase, nutrients become limited after consumption in the exponential<br />

phase <strong>of</strong> growth, which is the trigger for the so-called stringent response. The stringent response<br />

comprises an adaptation <strong>of</strong> the bacteria’s metabolism to situations <strong>of</strong> nutrient limitation or<br />

starvation, which includes induction <strong>of</strong> stress resistance, decelerated growth and alleviated<br />

metabolism. Many <strong>of</strong> the necessary changes are mediated <strong>by</strong> transcriptional variation (Condon et<br />

197


Maren Depke<br />

Discussion and Conclusions<br />

al. 1995, Crosse et al. 2000, Anderson KL et al. 2006, Wolz et al. 2010). In a comparison between<br />

the stationary phase gene expression pr<strong>of</strong>ile and that <strong>of</strong> the late serum/CO 2 control (each in<br />

relation to exponential phase growth) a high overlap was recognized. This included reduced<br />

expression <strong>of</strong> ribosomal protein genes, tRNA synthetase genes, translation elongation factor<br />

genes and a clear trend <strong>of</strong> induction <strong>of</strong> the stringent response regulator relA as it was reported in<br />

literature (Anderson KL et al. 2006). Thus, the gene expression in the late serum/CO 2 controls<br />

resembled the stationary phase/stringent response. A corresponding similarity was not observed<br />

for the internalized staphylococci as well as for the early control samples. In conclusion from<br />

these observations, it was decided that the 1 h serum/CO 2 control sample, although not timematched,<br />

was the most appropriate baseline sample for this infection experiment study.<br />

Since the shift <strong>of</strong> bacterial cells from aerobic agitated culture to infection settings in cell<br />

culture plates and 5 % CO 2 atmosphere probably resulted in a change <strong>of</strong> oxygen availability, genes<br />

harboring binding sites for Rex, the central anaerobiosis regulator and transcriptional repressor<br />

(Pagels et al. 2010), were analyzed to elucidate whether the changes <strong>of</strong> oxygen availability lead<br />

to an anaerobiosis gene expression signature (this study). De-repression <strong>of</strong> Rex-binding site<br />

containing genes was visible during anaerobic incubation, but despite the shift from agitated<br />

culture to cell culture plate, internalized samples did not show the same pattern <strong>of</strong> anaerobic<br />

gene expression. Contrarily, many <strong>of</strong> these Rex binding sites containing genes, which were in<br />

trend or significantly induced in anaerobically incubated samples (indicating de-repression after<br />

Rex lost its DNA affinity in anaerobic conditions), were in trend or significantly repressed in<br />

internalized staphylococci in S9 cells (indicating sustained DNA binding activity <strong>of</strong> Rex). Thus, it<br />

was concluded that internalized staphylococci did not suffer from reduced oxygen availability.<br />

One <strong>of</strong> the first observations during the analysis <strong>of</strong> the internalized staphylococcal gene<br />

expression pattern was that the response regulator gene saeR <strong>of</strong> the two-component system<br />

SaeRS was induced in internalized staphylococci 6.5 h after start <strong>of</strong> infection and that the sensor<br />

component gene saeS was significantly differentially expressed but only did not exceed a fold<br />

change <strong>of</strong> 2. This observation led to a more detailed analysis <strong>of</strong> the SaeRS regulon in internalized<br />

staphylococci.<br />

In 2006, Rogasch and coworkers published a transcriptomic and proteomic <strong>characterization</strong><br />

study <strong>of</strong> the SaeRS regulon (Rogasch et al. 2006). Of the genes defined to be regulated in that<br />

publication, 21 were found to be regulated (mostly increased) in at least one <strong>of</strong> the two analyzed<br />

time points <strong>of</strong> internalized staphylococci in the study described in this thesis. These included the<br />

auto-induced saeR, adhesins, serine protease, toxins, immune-evasive genes and others. Thus,<br />

most probably the SaeRS two-component system was activated in internalized staphylococci.<br />

The SaeRS system was furthermore studied <strong>by</strong> DNA array analysis using a saeS gene<br />

replacement mutant (Sa371ko) and its parental strain, the clinical isolate S. aureus WCUH29<br />

(Liang et al. 2006). Less than 20 genes (positive influence on coa, fnbB, fnb, efb, hla, hlb, hlgC,<br />

saeS, SA1000, which corresponds to SAOUHSC_01110, and others; negative influence on agrA<br />

and a gene coding for a hypothetical protein) were observed to be SaeRS dependently regulated.<br />

In vivo, the saeS mutant strain resulted in less bacterial load in the kidney compared to the wild<br />

type strain in a hematogenous pyelonephritis model <strong>of</strong> i. v. infected mice (Liang et al. 2006).<br />

Interestingly, S. aureus deficient for SaeS exhibited less adhesion to and less invasion in human<br />

lung epithelial A549 cells and resulted in a reduced rate <strong>of</strong> <strong>host</strong> cell apoptosis which was possibly<br />

mediated in parts <strong>by</strong> reduced expression <strong>of</strong> hla (Liang et al. 2006). Furthermore, negative<br />

influence on adhesion and internalization was documented for efb and SA1000 replacement<br />

198


Maren Depke<br />

Discussion and Conclusions<br />

mutants. Thus, especially saeRS as regulatory system, but also efb and SA1000 (SAOUHSC_01110)<br />

contribute to adhesion and invasion (Liang et al. 2006). Strikingly, these genes were observed to<br />

be induced (saeR, efb) or in trend induced (saeS; SAOUHSC_01110; data not shown for<br />

SAOUHSC_01110) in internalized staphylococci in the study described in this thesis. Furthermore,<br />

also coa, fnbB, hla, and hlgC, determined as SaeRS-dependent <strong>by</strong> Liang and colleagues, were<br />

induced in internalized staphylococci in this study. The final conclusion <strong>of</strong> Liang et al. stated that<br />

“activation <strong>of</strong> the SaeRS system is required for S. aureus to adhere to and invade epithelial cells”.<br />

Supportingly, such activation was also concluded in this study.<br />

Nevertheless, the activation SaeRS and induction <strong>of</strong> saeRS might depend on further<br />

experimental aspects which are not defined yet. Infection and gene expression studies <strong>by</strong> Garzoni<br />

and coworkers did not reveal induction but repression <strong>of</strong> saeR and saeS in S. aureus 6850 upon<br />

internalization in human lung epithelial A549 cells (Garzoni et al. 2007). Contrarily, upon<br />

phagocytosis <strong>by</strong> human polymorphonuclear leukocytes the induction <strong>of</strong> saeR and saeS was<br />

observed in different S. aureus strains (Voyich et al. 2005).<br />

While adherence <strong>of</strong> sae mutant and wild type to endothelial cells did not differ, the mutant<br />

was less invasive than the wild type (Steinhuber et al. 2003). Further affirmation <strong>of</strong> the<br />

participation <strong>of</strong> saeS in infection models was derived from mutagenesis studies in combination<br />

with murine systemic infection. Here, non-functional saeS led to attenuated virulence (Benton et<br />

al. 2004). A different bacterial strain background led to similar observations <strong>of</strong> reduced virulence<br />

in mice (Rampone et al. 1996). Also Goerke et al. identified SaeRS as important regulator which is<br />

active in vivo, although the deletion mutant did not yield different bacterial densities compared<br />

to the wild type in a device-related infection model (Goerke et al. 2005).<br />

Besides the already mentioned saeR and saeS, three other regulators were observed with<br />

differential expression in internalized staphylococci (sarT, sarU, rot). Since sarT and sarU are less<br />

well characterized and since all these three genes were repressed, the effect <strong>of</strong> this differential<br />

expression cannot easily be rated in the setting <strong>of</strong> the in vitro infection model.<br />

Fittingly to the concluded activity <strong>of</strong> the SaeRS system in internalized staphylococci,<br />

membrane-bound adhesins, which are partly known to be regulated <strong>by</strong> SaeRS (Liang et al. 2006),<br />

were induced (fnbA, fnbB, clfA, clfB).<br />

Fibronectin binding proteins (FnBPs) are – together with other surface proteins – important<br />

mediators <strong>of</strong> bacterial adhesion to <strong>host</strong> cells. They exhibit an even more central position for<br />

internalization processes. FnBPs were partly required for adhesion and mainly required for<br />

internalization into the bovine mammary gland epithelial cell line, MAC-T (Dziewanowska et al.<br />

1999). The mechanism includes binding <strong>of</strong> FnBP to β 1 integrins via a bridge formed <strong>by</strong> fibronectin,<br />

but also direct binding <strong>of</strong> FnBP to <strong>host</strong> membrane-located Hsp60 (Dziewanowska et al. 2000).<br />

Anyway, another study demonstrated variability <strong>of</strong> fnb genes between different clinical isolates<br />

as well as variation in the ability to adhere to fibronectin <strong>by</strong> the different isolates. Strains which<br />

were derived from orthopaedic implant-associated infection showed higher adherence than the<br />

other isolates. Community-acquired invasive disease isolates harbored more <strong>of</strong>ten two fnb genes<br />

than carriage isolates, but the number <strong>of</strong> fnb genes correlated only to a low extent with the<br />

adhesion capacity (Peacock et al. 2000).<br />

Fibronectin-binding protein deficiency reduced the colonization <strong>of</strong> heart tissue in a rat<br />

endocarditis model indicating a role <strong>of</strong> FnBP in heart valve infection. Other tissue samples did not<br />

show the same effect (Kuypers/Proctor 1989). In order to avoid compensation <strong>of</strong> a deleted gene<br />

<strong>by</strong> other, functionally redundant staphylococcal genes, an overexpression study <strong>of</strong> clfA and fnbA<br />

199


Maren Depke<br />

Discussion and Conclusions<br />

in Lactococcus lactis, a bacterium with low <strong>pathogen</strong>ic potential, was performed (Que et al.<br />

2001). Possession <strong>of</strong> ClfA or FnbA rendered L. lactis as adherent to fibrinogen or fibronectin as<br />

the gene-donor S. aureus strain. Furthermore, overexpressing L. lactis strains were able to cause<br />

endocarditis in a rat model with a similar minimal infection dose as S. aureus. The authors stated<br />

that this overexpression experiment proved the involvement <strong>of</strong> clfA and fnbA in endocarditis with<br />

higher confidence than deletion experiments (Que et al. 2001).<br />

Thus, fnb genes have proven to be relevant for adhesion, invasion and in vivo infection<br />

situations. Surprisingly, the induction <strong>of</strong> adhesins was observed after internalization <strong>of</strong> S. aureus<br />

RN1HG into S9 cells (this study). It can be speculated that this induction took place in advance to<br />

react to a possible re-liberation from the <strong>host</strong> cell e. g. after <strong>host</strong> cell death.<br />

In internalized staphylococci, also soluble adhesins were induced (eap, coa, vwb, emp, efb).<br />

These are also associated with other functions like immune-evasion. Eap, whose expression is<br />

essentially mediated <strong>by</strong> sae (Harraghy et al. 2005), is one <strong>of</strong> the examples for a staphylococcal<br />

protein which is related to a variety <strong>of</strong> functions (Harraghy et al. 2003). Eap was reported for<br />

example to mediate adhesion to cultured fibroblasts (Hussain et al. 2002), to enhance<br />

internalization into human fetal lung fibroblasts and HACAT keratinocytes (Haggar et al. 2003), to<br />

impair wound healing via inhibition <strong>of</strong> angiogenesis (Athanasopoulos et al. 2006), to block Ras<br />

activation and signal transduction cascade which leads to anti-angiogenesis (Sobke et al. 2006),<br />

to induce pro-inflammatory IL-6 and TNF-α in human peripheral blood mononuclear cells (Scriba<br />

et al. 2008), but also to act anti-inflammatorily <strong>by</strong> inhibition <strong>of</strong> leukocyte recruitment via<br />

blockade <strong>of</strong> ICAM1 and hence lead to immune evasion (Chavakis et al. 2002, Haggar et al. 2004).<br />

Thus, it was not surprising to find induced expression also after internalization into S9 cells (this<br />

study). Further immune-evasion related genes were increased in S9-internalized S. aureus<br />

RN1HG, e. g. chp, whose gene product CHIPS, chemotaxis inhibitory protein <strong>of</strong> staphylococci, is<br />

known to block the receptor for the chemotactic complement fragment C5a and for bacterialderived<br />

formylated peptides (de Haas et al. 2004, Murdoch/Finn 2000). Also induced efb, coding<br />

for extracellular fibrinogen-binding protein Efb, interferes with complement action, i. e.<br />

opsonization, <strong>by</strong> inhibition <strong>of</strong> C3b-binding to bacterial surfaces (Lee et al. 2004a, 2004b). Thus,<br />

internalized staphylococci induce certain genes which would benefit their survival in an in vivo<br />

situation.<br />

In aerobically living cells, reactive oxygen intermediates like hydroxyl radicals (OH•), hydroxide<br />

(OH – ), superoxide anion (O – 2 ), or hydrogen peroxide (H 2 O 2 ) occur in normal metabolism and may<br />

damage cellular structures like DNA (Fridovich 1978, Imlay/Linn 1988). Phagocytes even produce<br />

high amounts <strong>of</strong> reactive oxygen intermediates during respiratory burst, e. g. superoxide anion <strong>by</strong><br />

their NADPH-oxidase, as part <strong>of</strong> their defense and killing strategy (Robinson 2009). S. aureus<br />

owns two genes coding for superoxide dismutases: sodA and sodM (Clements et al. 1999,<br />

Valderas/Hart 2001). The enzyme participates in the detoxification <strong>of</strong> reactive oxygen<br />

intermediates <strong>by</strong> catalyzing the reaction <strong>of</strong> two superoxide anion molecules to oxygen and<br />

hydrogen peroxide, <strong>of</strong> which two molecules are consequently decomposed to oxygen and water<br />

<strong>by</strong> catalase. The complete enzyme constitutes <strong>of</strong> 2 homo- or heterodimeric subunits. The gene<br />

sodM is characteristic for S. aureus since coagulase-negative staphylococci like S. epidermidis do<br />

not own this gene or a homologue (Valderas et al. 2002).<br />

In the presence <strong>of</strong> manganese, the internal (methyl viologen) and external (xanthine/xanthine<br />

oxidase) generation <strong>of</strong> O –<br />

2 in S. aureus led to increased expression <strong>of</strong> sodA and sodM,<br />

respectively, but not without the addition <strong>of</strong> manganese (Karavolos et al. 2003). The observation<br />

200


Maren Depke<br />

Discussion and Conclusions<br />

<strong>of</strong> increased sodA and repressed sodM in internalized staphylococci (this study) fits to an<br />

independent regulation <strong>of</strong> both genes. Anyway, the regulation <strong>of</strong> staphylococcal sod genes has<br />

not been intensively studied until now. The analysis <strong>of</strong> mutants in the regulator genes sigB, agr,<br />

or sarA did not influence the sodA expression in experiments <strong>of</strong> Clements et al. in 1999. Use <strong>of</strong> a<br />

sigB deficient S. aureus strain gave hints for a negative influence <strong>of</strong> σ B on the expression <strong>of</strong> sodM<br />

(Karavolos et al. 2003). Contradictory to the results <strong>of</strong> Clements and colleagues, SarA was<br />

reported as negative regulator <strong>of</strong> especially sodM but also sodA expression via its activity as<br />

transcriptional repressor since gel-shift analysis and consensus sequence search revealed SarAbinding<br />

to the sod promoter regions (Ballal/Manna 2009). The observations <strong>of</strong> regulatory<br />

influence do not completely explain the expression pattern in this study, and further attempts <strong>of</strong><br />

explanation would be too speculative because <strong>of</strong> only little existing knowledge on the<br />

internalized gene expression regulation.<br />

Detoxification <strong>of</strong> reactive oxygen intermediates and correlated gene expression might have<br />

impact on virulence since this detoxification, i. e. evasion <strong>of</strong> immune defense, can give an<br />

advantage to the bacterial cell. Studies from literature give contradictory evidence. Superoxide<br />

dismutase did not correlate with lethality in mouse infection but catalase was proposed to<br />

enhance virulence using clinical isolates and Wood-46 strain <strong>of</strong> S. aureus (Mandell 1975).<br />

S. aureus 8325-4 sodA mutant did not possess reduced virulence in comparison to wild type<br />

strains in a mouse abscess model (Clements et al. 1999). S. aureus SH1000 sodA, sodM, and<br />

sodA sodM single and double mutants exhibited reduced virulence in a murine model <strong>of</strong><br />

subcutaneous infection (Karavolos et al. 2003). The SigB – phenotype <strong>of</strong> 8325-4, which has been<br />

reversed in SH1000 (Kullik/Giachino 1997, Horsburgh et al. 2002b), might have contributed to<br />

these observations. The superoxide dismutase gene sodA was reported to be linked to the<br />

bacterial acid-adaptive response (Clements et al. 1999). Such function might provide a link to the<br />

expression pattern in internalized staphylococci, which probably encounter reduced pH in the<br />

phagosome/endo-lysosome.<br />

Internalized staphylococci induced two pairs <strong>of</strong> bicomponent toxins, lukD/lukE, hlgB/hlgC, and<br />

alpha-hemolysin, hla. The subunits <strong>of</strong> the bicomponent toxins were described to be<br />

interchangeable and to result in different toxins with distinct properties, which led to the<br />

activation <strong>of</strong> granulocytes (König et al. 1997). Leukocidins target the membrane and lead to<br />

disruption <strong>of</strong> <strong>host</strong> cells for example <strong>of</strong> human polymorphonuclear neutrophils <strong>by</strong> PVL (Colin et al.<br />

1994) or <strong>of</strong> erythrocytes <strong>by</strong> a LukF/Hlg combination (Nguyen et al. 2003) and cause severe<br />

inflammatory damage in a model <strong>of</strong> toxin injection into rabbit eyes (Siqueira et al. 1997).<br />

LukD/LukE were first characterized <strong>by</strong> Gravet et al. in 1998. From a set <strong>of</strong> 429 S. aureus isolates,<br />

the prevalence <strong>of</strong> lukD/lukE was determined as 82 % in blood and 60.5 % in nasal isolates,<br />

whereas hlg was detected in almost 100 % <strong>of</strong> isolates (von Eiff et al 2004). The LukD/LukE<br />

combination was shown to result in dermonecrosis after infection <strong>of</strong> rabbit skin. The<br />

dermonecrosis inducing potency <strong>of</strong> LukD or LukE was reduced when combinations with other<br />

toxin subunits were applied. Especially the combinations with HlgB led to reduced activity.<br />

HlgB/HlgC possessed lytic activity for erythrocytes, but the LukD/LukE combination did not lead<br />

to erythrocyte lysis. Erythrocyte lysis could not be achieved <strong>by</strong> combining LukD or LukE with<br />

other toxin subunits. Finally, human polymorphonuclear leukocytes (granulocytes) could be<br />

activated <strong>by</strong> HlgB/HlgC, LukE/HlgB, LukD/HlgC, and LukD/LukE which decreased potency in the<br />

given order (Gravet et al. 1998). Thus, the different toxins have a distinct functional bias as well<br />

as potency. The specific effects on S9 cells or in vivo, e. g. in the lung, still need to be determined.<br />

201


Maren Depke<br />

Discussion and Conclusions<br />

Similarly, the expression <strong>of</strong> extracellular and membrane-bound enzymes, which contribute to<br />

virulence or spreading, was difficult to interpret. Expression <strong>of</strong> serine proteases (splABCDEF) and<br />

lipase (lip) was increased, whereas the expression <strong>of</strong> hyaluronate lyase (hysA), a membrane<br />

bound serine protease (htrA), nuclease (nuc), and glutamyl endopeptidase alias V8 peptidase<br />

(sspA) was repressed in internalized staphylococci. Some spl genes were found to be SaeRSregulated<br />

(Rogasch et al. 2006). Thus, the induction <strong>of</strong> the spl operon fits to the proposed activity<br />

<strong>of</strong> the SaeRS two-component system.<br />

Gene expression changes which might have influence on cell wall remodeling were also<br />

observed in internalized staphylococci. Surprisingly, this included repression <strong>of</strong> the dlt operon.<br />

The proteins encoded <strong>by</strong> dlt participate in incorporation and esterification <strong>of</strong> D-alanine residues<br />

into the cell wall teichoic acids. In this way, the negative charge <strong>of</strong> the cell wall structures is<br />

reduced and thus is less attractant for positively charged antimicrobial molecules (Peschel et al.<br />

1999). With the aim to evade the immune response, an induction <strong>of</strong> dlt expression had been<br />

expected. In parallel, the induction <strong>of</strong> peptidoglycan hydrolase lytM and staphylococcal secretory<br />

antigen ssaA was recorded (this study), which are also involved in cell wall remodeling (Dubrac et<br />

al. 2007). Repression <strong>of</strong> dlt and induction <strong>of</strong> lytM were also observed <strong>by</strong> Garzoni and coworkers<br />

(Garzoni et al. 2007). Furthermore, the induction <strong>of</strong> prsA, coding for a foldase involved in<br />

translocation and folding <strong>of</strong> secreted proteins (Sarvas et al. 2004), was observed in internalized<br />

staphylococci 2.5 h after infection <strong>of</strong> S9 cells (this study). This regulation was in accordance with<br />

protein abundance data in internalized staphylococci which were determined <strong>by</strong> Sandra Scharf<br />

(Schmidt et al. 2010). She also found an increased abundance <strong>of</strong> VraR, regulatory component <strong>of</strong><br />

the VraRS two-component system and known to be positively auto-regulated (Kuroda et al.<br />

2003), which led to the examination <strong>of</strong> vraRS gene expression: vraR (1.8) and vraS (1.6) were not<br />

significantly different in internalized staphylococci (2.5 h) in comparison to control, but the fold<br />

change values indicated a slight trend for induction. Hence, the observed differential expression<br />

<strong>of</strong> cell wall remodeling related genes was in agreement with the proteome data <strong>of</strong> Sandra Scharf.<br />

Additionally, genes described to belong to the VraRS regulon (Kuroda et al. 2003) were examined.<br />

Internalized staphylococci induced about 20 % <strong>of</strong> the published VraRS regulon (Kuroda et al.<br />

2003) at the same time point when a trend <strong>of</strong> vraRS induction was detected. Since the VraRS<br />

two-component system was suggested to be activated upon cell wall damage or inhibition <strong>of</strong> cell<br />

wall biosynthesis (Kuroda et al. 2000, Kuroda et al. 2003), the observation <strong>of</strong> cell wall remodeling<br />

associated gene expression changes might indicate a reaction to stressful influences which<br />

possibly operate on the bacterial cell wall after internalization into the <strong>host</strong> cell. Another possible<br />

explanation proposed <strong>by</strong> Garzoni and colleagues (2007) is that the gene induction (e. g. lytM) is<br />

associated with cell division processes.<br />

The gene expression analysis <strong>of</strong> internalized staphylococci in S9 cells led to the recognition <strong>of</strong><br />

changes in the expression <strong>of</strong> metabolic genes. Purine biosynthesis, tRNA synthetases, and<br />

glycolysis were repressed, whereas gluconeogenesis, TCA cycle enzyme genes, and especially<br />

genes related to amino acid biosynthesis were induced. The induction <strong>of</strong> amino acid biosynthesis<br />

and TCA cycle genes as well as the repression <strong>of</strong> purine biosynthesis and tRNA synthetase genes<br />

was clearly confirmed in the proteome analysis <strong>of</strong> Sandra Scharf. Additionally, transporter genes<br />

were differentially expressed. Induction was observed especially for phosphate transporters, but<br />

also for transporters <strong>of</strong> phosphonate, methionine, peptide/oligopeptide, sugar/maltose,<br />

osmoprotectants, amino acids, and gluconate. Repression was detected for urea,<br />

202


Maren Depke<br />

Discussion and Conclusions<br />

spermidine/putrescine, sodium/glutamate, arginine/ornithine, glycine betaine, amino acid,<br />

choline, citrate, xanthine, fructose, and lactate transporters.<br />

A good metabolic performance, especially the unaffected ability for amino acid biosynthesis<br />

and for the uptake <strong>of</strong> compounds, is important for the survival <strong>of</strong> <strong>pathogen</strong>s and has been<br />

reported to be linked not only to auxotrophies, but also to virulence. In a mutagenesis study,<br />

which was combined with a model <strong>of</strong> murine systemic infection, 24 genes were identified whose<br />

mutation resulted in attenuation <strong>of</strong> S. aureus virulence. Of these, 9 genes (38 %) were part <strong>of</strong><br />

purine or amino acid biosynthesis, and 6 genes (25 %) coded for membrane transporters (Benton<br />

et al. 2004). Of these virulence associated genes, four genes related to amino acid biosynthesis<br />

(lysC, dapB, trpF, asd) and pstS, phosphate transporter, were induced in internalized<br />

staphylococci at least in one <strong>of</strong> the two analyzed time points (this study). Additionally, amino acid<br />

biosynthesis genes tyrA and pycA were induced in trend with a significant p-value but a fold<br />

change value <strong>of</strong> less than the cut<strong>of</strong>f.<br />

Few gene expression studies <strong>of</strong> internalized staphylococci are published until now. One <strong>of</strong><br />

them has already been cited above in selected details. Garzoni and coworkers analyzed<br />

internalized staphylococci in epithelial cells using a microarray approach (Garzoni et al. 2007).<br />

Although the general setting resembles that applied in this study, important differences in the<br />

experimental procedures as well as in the results can be found between both studies. Most<br />

importantly, staphylococcal strain (S. aureus 6850 vs. RN1HG) as well as <strong>host</strong> cell line (A549 vs.<br />

S9) differed. Garzoni et al. mention exactly this topic and conclude that “certain combinations <strong>of</strong><br />

<strong>host</strong> cells and bacteria can result in different biological outcomes”, which is clearly supported in<br />

the comparison <strong>of</strong> their and this study. Furthermore, infection took place in a cell culture<br />

medium without (Garzoni et al. 2007) or with serum supplement (this study), with washed<br />

(Garzoni et al. 2007) or non-washed, exoproteins containing (this study) bacterial suspensions, in<br />

a multiplicity <strong>of</strong> infection (MOI) <strong>of</strong> 100 for 30 min (Garzoni et al. 2007) or in a MOI <strong>of</strong> 25 for 1 h<br />

(this study). Also RNA preparation and sample processing methods differed: depletion <strong>of</strong><br />

eukaryotic RNA and amplification step (Garzoni et al. 2007) or utilization <strong>of</strong> RNA without<br />

depletion and amplification since pure bacterial RNA <strong>of</strong> high quality was available in sufficient<br />

amounts (this study). Finally, different arrays were used. Both studies approximately match in the<br />

analyzed time points <strong>of</strong> 2 h/2.5 h and 6 h/6.5 h after infection.<br />

The total numbers <strong>of</strong> differentially expressed genes were higher in the literature reference<br />

with 1042 (Garzoni et al. 2007) and 565 (this study, annotated genes) for the early time point and<br />

766 (Garzoni et al. 2007) and 489 (this study, annotated genes) for the later time point, and<br />

Garzoni et al. detected a higher fraction <strong>of</strong> repressed genes whereas this study resulted in a<br />

higher number <strong>of</strong> induced genes. Similar in both studies, comparisons between non-adherent<br />

and mock infection samples (staphylococci in infection medium) did not result in the detection <strong>of</strong><br />

differential gene expression. In a very rough comparison, at least 100 genes were differentially<br />

expressed between internalized and control staphylococci in both studies. Garzoni and<br />

colleagues describe differential expression <strong>of</strong> genes or functional groups <strong>of</strong> genes, which were<br />

recognized also in this study, e. g. repression <strong>of</strong> tRNA synthetase genes, induction <strong>of</strong> fnbAB,<br />

induction <strong>of</strong> hlgB and lukE, induction <strong>of</strong> sodA, repression <strong>of</strong> sspA, induction <strong>of</strong> lytM, and induction<br />

<strong>of</strong> transporters. Different results were revealed, e. g. clfB repressed (Garzoni et al. 2007) or<br />

induced (this study), hla not differentially expressed (Garzoni et al. 2007) or induced (this study),<br />

cap operon repressed (Garzoni et al. 2007) or not differentially expressed (this study), TCA cycle<br />

203


Maren Depke<br />

Discussion and Conclusions<br />

genes repressed (Garzoni et al. 2007) or induced (this study), saeRS repressed (Garzoni et al.<br />

2007) or induced (this study).<br />

Clearly, the comparability is limited due to experimental differences, which underlines the<br />

importance <strong>of</strong> performing global molecular studies (transcriptome and proteome pr<strong>of</strong>iling) in<br />

different models to finally achieve a complete picture <strong>of</strong> the characteristics <strong>of</strong> <strong>host</strong>-<strong>pathogen</strong><br />

<strong>interactions</strong>.<br />

Such different models include e. g. other epithelial cell lines, like the S9 cells used in this study,<br />

but can also be extended to immune cells. Such a study was performed <strong>by</strong> Voyich and colleagues<br />

in 2005. The authors analyzed the gene expression pr<strong>of</strong>ile <strong>of</strong> different S. aureus strains upon<br />

phagocytosis <strong>by</strong> human neutrophils (Voyich et al. 2005). Again, several experimental parameters<br />

differed between the studies <strong>of</strong> Voyich and colleagues (2005), Garzoni and colleagues (2007), and<br />

this study, partly resulting from the very different <strong>host</strong> cell type employed <strong>by</strong> Voyich and<br />

coworkers. The authors described immune evasion mechanisms <strong>of</strong> S. aureus, which are initiated<br />

<strong>by</strong> transcriptional changes. First, the authors observed induction <strong>of</strong> stress response genes, e. g.<br />

superoxide dismutases sodA and sodM and several other genes involved in counteracting the<br />

influence <strong>of</strong> reactive oxygen intermediates. An as strong response was not observed in this study,<br />

which can be explained <strong>by</strong> the innate function <strong>of</strong> neutrophils to kill <strong>pathogen</strong>s, which is not given<br />

for epithelial cells. Thus, staphylococci internalized in S9 cells probably encounter less harmful<br />

molecules than staphylococci after uptake <strong>by</strong> neutrophils. The induction <strong>of</strong> TCA cycle genes after<br />

phagocytosis <strong>by</strong> neutrophils (Voyich et al. 2005) corresponds with results from this study using S9<br />

cells. Voyich and coworkers observed repression <strong>of</strong> genes involved in cell envelope synthesis, cell<br />

division, and replication, which was not as obvious in S9 cell internalized staphylococci.<br />

Nevertheless, cell wall remodeling processes were concluded from gene expression data (this<br />

study). Furthermore, staphylococci in neutrophils induced several toxins and adhesins (Voyich et<br />

al. 2005), <strong>of</strong> which some were also observed after S9 cell internalization, e. g. hlgB, hlgC, lukD,<br />

lukE; fnbB, coa, clfA (this study). The toxin gene induction appeared to be stronger in the<br />

neutrophil than in the S9 study, although hla was induced after internalization in S9 cells and not<br />

after phagocytosis <strong>by</strong> neutrophils. Anyway, the different staphylococcal strains, which were used<br />

in the studies, contribute considerably to the characteristics <strong>of</strong> the gene expression signature<br />

(Voyich et al. 2005). In the neutrophil phagocytosis dependent gene expression pattern,<br />

differential expression <strong>of</strong> regulatory genes was reported (Voyich et al. 2005). These included<br />

increased expression <strong>of</strong> vraRS, saeRS, and sarA. While induction or a trend <strong>of</strong> induction <strong>of</strong> saeRS<br />

and vraRS was also visible in this study, sarA was not differentially expressed when staphylococci<br />

were internalized in S9 cells. Furthermore, repression <strong>of</strong> agr or sigB was not observed in this<br />

study, but in the study <strong>of</strong> Voyich and coworkers.<br />

In total, after comparison <strong>of</strong> the results <strong>of</strong> Voyich and colleagues (2005), Garzoni and<br />

colleagues (2007), and this study, it becomes clear that similarities in certain aspects do not<br />

necessarily lead to completely consistent results. This probably depends to a large extent on the<br />

combination <strong>of</strong> <strong>host</strong> cell and bacterial strain, but also on further experimental conditions which<br />

varied strongly between the studies. Nevertheless, parts <strong>of</strong> the results were also found in other<br />

studies. Since an admittedly slightly imprecise attempt to specify real internalization specific gene<br />

expression that did not occur in control samples revealed virulence associated genes in the study<br />

described in this thesis, the recorded gene expression pattern is probably physiologically<br />

relevant. Therefore, the gene expression pr<strong>of</strong>iling turned out to be an important basic study<br />

which will entail follow-up experiments like studies <strong>of</strong> bacterial mutants, comparison <strong>of</strong> different<br />

204


Maren Depke<br />

Discussion and Conclusions<br />

<strong>host</strong> cell lines, but also confirmation and validation experiments <strong>by</strong> application <strong>of</strong> different<br />

techniques like quantitative real-time RT-PCR. Of course, the broadening <strong>of</strong> experiments to<br />

in vivo studies would be most interesting. Here, the most prominent problems are difficulties in<br />

recovery <strong>of</strong> bacterial cells from tissue and the resulting low amount <strong>of</strong> sample or the sample<br />

contamination with <strong>host</strong> material. In this setting, application <strong>of</strong> quantitative real-time RT-PCR<br />

probably will be the method <strong>of</strong> choice, which has already proven to successfully help resolving<br />

colonization expression patterns (Burian et al. 2010a, 2010b).<br />

Until now, the tiling array data have been used as expression data set under restriction to<br />

already annotated genomic content <strong>of</strong> S. aureus. The analysis is not yet finished. Although new<br />

transcribed units have been defined for which interesting regulation patterns in the infection<br />

experiments were visible, the sequences will have to be categorized and validated.<br />

Internalization dependently regulated sequences later will be targets for further <strong>characterization</strong><br />

experiments. It will be an interesting task to compare the resulting information on new genes<br />

with that generated with the competing approach <strong>of</strong> high-throughput transcriptome sequencing<br />

(Beaume et al 2010a, 2010b).<br />

205


Maren Depke<br />

R E F E R E N C E S<br />

Adams JM.<br />

Ways <strong>of</strong> dying: multiple pathways to apoptosis. Genes Dev. 2003 Oct 15;17(20):2481-95.<br />

Aikawa T, Matsutaka H, Takezawa K, Ishikawa E.<br />

Gluconeogenesis and amino acid metabolism. I. Comparison <strong>of</strong> various precursors for hepatic gluconeogenesis<br />

in vivo. Biochim Biophys Acta. 1972 Sep 15;279(2):234-44.<br />

Akhtar MK, Kelly SL, Kaderbhai MA.<br />

Cytochrome b 5 modulation <strong>of</strong> 17α hydroxylase and 17-20 lyase (CYP17) activities in steroidogenesis. J Endocrinol.<br />

2005 Nov;187(2):267-74.<br />

Aki M, Shimbara N, Takashina M, Akiyama K, Kagawa S, Tamura T, Tanahashi N, Yoshimura T, Tanaka K, Ichihara A.<br />

Interferon-γ induces different subunit organizations and functional diversity <strong>of</strong> proteasomes. J Biochem. 1994 Feb;<br />

115(2):257-69.<br />

Alberati-Giani D, Ricciardi-Castagnoli P, Köhler C, Cesura AM.<br />

Regulation <strong>of</strong> the kynurenine metabolic pathway <strong>by</strong> interferon-γ in murine cloned macrophages and microglial cells.<br />

J Neurochem. 1996 Mar;66(3):996-1004.<br />

Alberati-Giani D, Malherbe P, Ricciardi-Castagnoli P, Köhler C, Denis-Donini S, Cesura AM.<br />

Differential regulation <strong>of</strong> indoleamine 2,3-dioxygenase expression <strong>by</strong> nitric oxide and inflammatory mediators in<br />

IFN-γ-activated murine macrophages and microglial cells. J Immunol. 1997 Jul 1;159(1):419-26.<br />

Alberda C, Graf A, McCargar L.<br />

Malnutrition: etiology, consequences, and assessment <strong>of</strong> a patient at risk. Best Pract Res Clin Gastroenterol.<br />

2006;20(3):419-39.<br />

Alberti KG, Eckel RH, Grundy SM, Zimmet PZ, Cleeman JI, Donato KA, Fruchart JC, James WP, Loria CM, Smith SC Jr;<br />

International Diabetes Federation Task Force on Epidemiology and Prevention; Hational Heart, Lung, and Blood<br />

Institute; American Heart Association; World Heart Federation; International Atherosclerosis Society; International<br />

Association for the Study <strong>of</strong> Obesity.<br />

Harmonizing the metabolic syndrome: a joint interim statement <strong>of</strong> the International Diabetes Federation Task Force<br />

on Epidemiology and Prevention; National Heart, Lung, and Blood Institute; American Heart Association; World<br />

Heart Federation; International Atherosclerosis Society; and International Association for the Study <strong>of</strong> Obesity.<br />

Circulation. 2009 Oct 20;120(16):1640-5. Epub 2009 Oct 5.<br />

Almeida RA, Matthews KR, Cifrian E, Guidry AJ, Oliver SP.<br />

Staphylococcus aureus invasion <strong>of</strong> bovine mammary epithelial cells. J Dairy Sci. 1996 Jun;79(6):1021-6.<br />

Amasheh S, Meiri N, Gitter AH, Schöneberg T, Mankertz J, Schulzke JD, Fromm M.<br />

Claudin-2 expression induces cation-selective channels in tight junctions <strong>of</strong> epithelial cells. J Cell Sci. 2002 Dec 15;<br />

115(Pt 24):4969-76.<br />

Andersen MH, Schrama D, thor Straten P, Becker JC.<br />

Cytotoxic T cells. J Invest Dermatol. 2006 Jan;126(1):32-41.<br />

Anderson KL, Roberts C, Disz T, Vonstein V, Hwang K, Overbeek R, Olson PD, Projan SJ, Dunman PM.<br />

Characterization <strong>of</strong> the Staphylococcus aureus heat shock, cold shock, stringent, and SOS responses and their effects<br />

on log-phase mRNA turnover. J Bacteriol. 2006 Oct;188(19):6739-56.<br />

Anderson SL, Carton JM, Lou J, Xing L, Rubin BY.<br />

Interferon-induced guanylate binding protein-1 (GBP-1) mediates an antiviral effect against vesicular stomatitis virus<br />

and encephalomyocarditis virus. Virology. 1999 Mar 30;256(1):8-14.<br />

Arnalich F, López J, Codoceo R, Jim nez M, Madero R, Montiel C.<br />

Relationship <strong>of</strong> plasma leptin to plasma cytokines and human survivalin sepsis and septic shock. J Infect Dis. 1999<br />

Sep;180(3):908-11.<br />

207


Maren Depke<br />

References<br />

Athanasopoulos AN, Economopoulou M, Orlova VV, Sobke A, Schneider D, Weber H, Augustin HG, Eming SA,<br />

Schubert U, Linn T, Nawroth PP, Hussain M, Hammes HP, Herrmann M, Preissner KT, Chavakis T.<br />

The extracellular adherence protein (Eap) <strong>of</strong> Staphylococcus aureus inhibits wound healing <strong>by</strong> interfering with <strong>host</strong><br />

defense and repair mechanisms. Blood. 2006 Apr 1;107(7):2720-7. Epub 2005 Nov 29.<br />

Autenrieth IB, Beer M, Bohn E, Kaufmann SH, Heesemann J.<br />

Immune responses to Yersinia enterocolitica in susceptible BALB/c and resistant C57BL/6 mice: an essential role for<br />

gamma interferon. Infect Immun. 1994 Jun;62(6):2590-9.<br />

Ayroldi E, Zollo O, Bastianelli A, Marchetti C, Agostini M, Di Virgilio R, Riccardi C.<br />

GILZ mediates the antiproliferative activity <strong>of</strong> glucocorticoids <strong>by</strong> negative regulation <strong>of</strong> Ras signaling. J Clin Invest.<br />

2007 Jun;117(6):1605-15. Epub 2007 May 10.<br />

Bag<strong>by</strong> GJ, Lang CH, Skrepnik N, Spitzer JJ.<br />

Attenuation <strong>of</strong> glucose metabolic changes resulting from TNF-α administration <strong>by</strong> adrenergic blockade. Am J<br />

Physiol. 1992 Apr;262(4 Pt 2):R628-35.<br />

Ballal A, Manna AC.<br />

Regulation <strong>of</strong> superoxide dismutase (sod) genes <strong>by</strong> SarA in Staphylococcus aureus. J Bacteriol. 2009 May;191(10):<br />

3301-10. Epub 2009 Mar 13.<br />

Banerjee RR, Rangwala SM, Shapiro JS, Rich AS, Rhoades B, Qi Y, Wang J, Rajala MW, Pocai A, Scherer PE, Steppan CM,<br />

Ahima RS, Obici S, Rossetti L, Lazar MA.<br />

Regulation <strong>of</strong> fasted blood glucose <strong>by</strong> resistin. Science. 2004 Feb 20;303(5661):1195-8.<br />

Bange FC, Flohr T, Buwitt U, Böttger EC.<br />

An interferon-induced protein with release factor activity is a tryptophanyl-tRNA synthetase. FEBS Lett. 1992 Mar<br />

30;300(2):162-6.<br />

Barter PJ, Nicholls S, Rye KA, Anantharamaiah GM, Navab M, Fogelman AM.<br />

Antiinflammatory properties <strong>of</strong> HDL. Circ Res. 2004 Oct 15;95(8):764-72.<br />

Bartolomucci A.<br />

Social stress, immune functions and disease in rodents. Front Neuroendocrinol. 2007 Apr;28(1):28-49. Epub 2007<br />

Feb 16.<br />

Barton LF, Cruz M, Rangwala R, Deepe GS Jr, Monaco JJ.<br />

Regulation <strong>of</strong> immunoproteasome subunit expression in vivo following <strong>pathogen</strong>ic fungal infection. J Immunol. 2002<br />

Sep 15;169(6):3046-52.<br />

Bateman BT, Donegan NP, Jarry TM, Palma M, Cheung AL.<br />

Evaluation <strong>of</strong> a tetracycline-inducible promoter in Staphylococcus aureus in vitro and in vivo and its application in<br />

demonstrating the role <strong>of</strong> sigB in microcolony formation. Infect Immun. 2001 Dec;69(12):7851-7.<br />

Baughman G, Wiederrecht GJ, Campbell NF, Martin MM, Bourgeois S.<br />

FKBP51, a novel T-cell-specific immunophilin capable <strong>of</strong> calcineurin inhibition. Mol Cell Biol. 1995 Aug;15(8):4395-<br />

402.<br />

Beale RJ, Bryg DJ, Bihari DJ.<br />

Immunonutrition in the critically ill: a systematic review <strong>of</strong> clinical outcome. Crit Care Med. 1999 Dec;27(12):2799-<br />

805.<br />

Beaume M, Hernandez D, Francois P, Schrenzel J.<br />

New approaches for functional genomic studies in staphylococci. Int J Med Microbiol. 2010 Feb;300(2-3):88-97.<br />

Epub 2009 Dec 14.<br />

Beaume M, Hernandez D, Farinelli L, Deluen C, Linder P, Gaspin C, Rom<strong>by</strong> P, Schrenzel J, Francois P.<br />

Cartography <strong>of</strong> methicillin-resistant S. aureus transcripts: detection, orientation and temporal expression during<br />

growth phase and stress conditions. PLoS One. 2010 May 20;5(5):e10725.<br />

Bekker LG, Freeman S, Murray PJ, Ryffel B, Kaplan G.<br />

TNF-α controls intracellular mycobacterial growth <strong>by</strong> both inducible nitric oxide synthase-dependent and inducible<br />

nitric oxide synthase-independent pathways. J Immunol. 2001 Jun 1;166(11):6728-34.<br />

Benoist C, Mathis D.<br />

Regulation <strong>of</strong> major histocompatibility complex class-II genes: X, Y and other letters <strong>of</strong> the alphabet. Annu Rev<br />

Immunol. 1990;8:681-715.<br />

Benton BM, Zhang JP, Bond S, Pope C, Christian T, Lee L, Winterberg KM, Schmid MB, Buysse JM.<br />

Large-scale identification <strong>of</strong> genes required for full virulence <strong>of</strong> Staphylococcus aureus. J Bacteriol. 2004 Dec;<br />

186(24):8478-89.<br />

208


Maren Depke<br />

References<br />

Bera A, Herbert S, Jakob A, Vollmer W, Götz F.<br />

Why are <strong>pathogen</strong>ic staphylococci so lysozyme resistant? The peptidoglycan O-acetyltransferase OatA is the major<br />

determinant for lysozyme resistance <strong>of</strong> Staphylococcus aureus. Mol Microbiol. 2005 Feb;55(3):778-87.<br />

Berrebi D, Bruscoli S, Cohen N, Foussat A, Migliorati G, Bouchet-Delbos L, Maillot MC, Portier A, Couderc J,<br />

Galanaud P, Peuchmaur M, Riccardi C, Emilie D.<br />

Synthesis <strong>of</strong> glucocorticoid-induced leucine zipper (GILZ) <strong>by</strong> macrophages: an anti-inflammatory and<br />

immunosuppressive mechanism shared <strong>by</strong> glucocorticoids and IL-10. Blood. 2003 Jan 15;101(2):729-38. Epub 2002<br />

Sep 12.<br />

Beutler B, Milsark IW, Cerami AC.<br />

Passive immunization against cachectin/tumor necrosis factor protects mice from lethal effect <strong>of</strong> endotoxin.<br />

Science. 1985 Aug 30;229(4716):869-71.<br />

reprint in: Classical article. J Immunol. 2008 Jul 1;181(1):7-9.<br />

Bisch<strong>of</strong>f M, Dunman P, Kormanec J, Macapagal D, Murphy E, Mounts W, Berger-Bächi B, Projan S.<br />

Microarray-based analysis <strong>of</strong> the Staphylococcus aureus σ B regulon. J Bacteriol. 2004 Jul;186(13):4085-99.<br />

Bjerketorp J, Nilsson M, Ljungh A, Flock JI, Jacobsson K, Frykberg L.<br />

A novel von Willebrand factor binding protein expressed <strong>by</strong> Staphylococcus aureus. Microbiology. 2002 Jul;<br />

148(Pt 7):2037-44.<br />

Blasi F, Sidenius N.<br />

The urokinase receptor: focused cell surface proteolysis, cell adhesion and signaling. FEBS Lett. 2010 May 3;584(9):<br />

1923-30. Epub 2009 Dec 27.<br />

Boasso A, Herbeuval JP, Hardy AW, Winkler C, Shearer GM.<br />

Regulation <strong>of</strong> indoleamine 2,3-dioxygenase and tryptophanyl-tRNA-synthetase <strong>by</strong> CTLA-4-Fc in human CD4 + T cells.<br />

Blood. 2005 Feb 15;105(4):1574-81. Epub 2004 Oct 5.<br />

Bodén MK, Flock JI.<br />

Fibrinogen-binding protein/clumping factor from Staphylococcus aureus. Infect Immun. 1989 Aug;57(8):2358-63.<br />

Boehm U, Klamp T, Groot M, Howard JC.<br />

Cellular responses to interferon-γ. Annu Rev Immunol. 1997;15:749-95.<br />

Bokarewa M, Tarkowski A.<br />

Human alpha-defensins neutralize fibrinolytic activity exerted <strong>by</strong> staphylokinase. Thromb Haemost. 2004 May;<br />

91(5):991-9.<br />

Bonventre JV.<br />

Phospholipase A2 and signal transduction. J Am Soc Nephrol. 1992 Aug;3(2):128-50.<br />

Bouillet P, Strasser A.<br />

BH3-only proteins - evolutionarily conserved proapoptotic Bcl-2 family members essential for initiating programmed<br />

cell death. J Cell Sci. 2002 Apr 15;115(Pt 8):1567-74.<br />

Breitbach K, Klocke S, Tschernig T, van Rooijen N, Baumann U, Steinmetz I.<br />

Role <strong>of</strong> inducible nitric oxide synthase and NADPH oxidase in early control <strong>of</strong> Burkholderia pseudomallei infection in<br />

mice. Infect Immun. 2006 Nov;74(11):6300-9. Epub 2006 Sep 25.<br />

Brosnan JT.<br />

Glutamate, at the interface between amino acid and carbohydrate metabolism. J Nutr. 2000 Apr;130(4S<br />

Suppl):988S-90S.<br />

Brucet M, Marqués L, Sebastián C, Lloberas J, Celada A.<br />

Regulation <strong>of</strong> murine Tap1 and Lmp2 genes in macrophages <strong>by</strong> interferon gamma is mediated <strong>by</strong> STAT1 and IRF-1.<br />

Genes Immun. 2004 Jan;5(1):26-35.<br />

Burian M, Wolz C, Goerke C.<br />

Regulatory adaptation <strong>of</strong> Staphylococcus aureus during nasal colonization <strong>of</strong> humans. PLoS One. 2010 Apr 6;<br />

5(4):e10040.<br />

Burian M, Rautenberg M, Kohler T, Fritz M, Krismer B, Unger C, H<strong>of</strong>fmann WH, Peschel A, Wolz C, Goerke C.<br />

Temporal expression <strong>of</strong> adhesion factors and activity <strong>of</strong> global regulators during establishment <strong>of</strong> Staphylococcus<br />

aureus nasal colonization. J Infect Dis. 2010 May 1;201(9):1414-21.<br />

Buwitt U, Flohr T, Böttger EC.<br />

Molecular cloning and <strong>characterization</strong> <strong>of</strong> an interferon induced human cDNA with sequence homology to a<br />

mammalian peptide chain release factor. EMBO J. 1992 Feb;11(2):489-96.<br />

209


Maren Depke<br />

References<br />

Byrne GI, Lehmann LK, Landry GJ.<br />

Induction <strong>of</strong> tryptophan catabolism is the mechanism for gamma-interferon-mediated inhibition <strong>of</strong> intracellular<br />

Chlamydia psittaci replication in T24 cells. Infect Immun. 1986 Aug;53(2):347-51.<br />

Campbell SJ, Hughes PM, Iredale JP, Wilcockson DC, Waters S, Docagne F, Perry VH, Anthony DC.<br />

CINC-1 is an acute-phase protein induced <strong>by</strong> focal brain injury causing leukocyte mobilization and liver injury.<br />

FASEB J. 2003 Jun;17(9):1168-70. Epub 2003 Apr 22.<br />

Capes SE, Hunt D, Malmberg K, Gerstein HC.<br />

Stress hyperglycaemia and increased risk <strong>of</strong> death after myocardial infarction in patients with and without diabetes:<br />

a systematic overview. Lancet. 2000 Mar 4;355(9206):773-8.<br />

Carlin JM, Borden EC, Sondel PM, Byrne GI.<br />

Interferon-induced indoleamine 2,3-dioxygenase activity in human mononuclear phagocytes. J Leukoc Biol. 1989<br />

Jan;45(1):29-34.<br />

Carter CC, Gorbacheva VY, Vestal DJ.<br />

Inhibition <strong>of</strong> VSV and EMCV replication <strong>by</strong> the interferon-induced GTPase, mGBP-2: differential requirement for<br />

wild-type GTP binding domain. Arch Virol. 2005 Jun;150(6):1213-20. Epub 2005 Feb 18.<br />

Casey R, Newcombe J, McFadden J, Bodman-Smith KB.<br />

The acute-phase reactant C-reactive protein binds to phosphorylcholine-expressing Neisseria meningitidis and<br />

increases uptake <strong>by</strong> human phagocytes. Infect Immun. 2008 Mar;76(3):1298-304. Epub 2008 Jan 14.<br />

Chambers HF.<br />

Methicillin resistance in staphylococci: molecular and biochemical basis and clinical implications. Clin Microbiol Rev.<br />

1997 Oct;10(4):781-91.<br />

Chan PF, Foster SJ, Ingham E, Clements MO.<br />

The Staphylococcus aureus alternative sigma factor σ B controls the environmental stress response but not starvation<br />

survival or <strong>pathogen</strong>icity in a mouse abscess model. J Bacteriol. 1998 Dec;180(23):6082-9.<br />

Chang SC, Momburg F, Bhutani N, Goldberg AL.<br />

The ER aminopeptidase, ERAP1, trims precursors to lengths <strong>of</strong> MHC class I peptides <strong>by</strong> a "molecular ruler"<br />

mechanism. Proc Natl Acad Sci U S A. 2005 Nov 22;102(47):17107-12. Epub 2005 Nov 14.<br />

Chatterjee-Kishore M, Kishore R, Hicklin DJ, Marincola FM, Ferrone S.<br />

Different requirements for signal transducer and activator <strong>of</strong> transcription 1α and interferon regulatory factor 1 in<br />

the regulation <strong>of</strong> low molecular mass polypeptide 2 and transporter associated with antigen processing 1 gene<br />

expression. J Biol Chem. 1998 Jun 26;273(26):16177-83.<br />

Chavakis T, Hussain M, Kanse SM, Peters G, Bretzel RG, Flock JI, Herrmann M, Preissner KT.<br />

Staphylococcus aureus extracellular adherence protein serves as anti-inflammatory factor <strong>by</strong> inhibiting the<br />

recruitment <strong>of</strong> <strong>host</strong> leukocytes. Nat Med. 2002 Jul;8(7):687-93. Epub 2002 Jun 24.<br />

Chen G, Goeddel DV.<br />

TNF-R1 signaling: a beautiful pathway. Science. 2002 May 31;296(5573):1634-5.<br />

Chen W, Suruga K, Nishimura N, Gouda T, Lam VN, Yokogoshi H.<br />

Comparative regulation <strong>of</strong> major enzymes in the bile acid biosynthesis pathway <strong>by</strong> cholesterol, cholate and taurine<br />

in mice and rats. Life Sci. 2005 Jul 1;77(7):746-57. Epub 2005 Mar 19.<br />

Cheung AL, Bayer AS, Zhang G, Gresham H, Xiong YQ.<br />

Regulation <strong>of</strong> virulence determinants in vitro and in vivo in Staphylococcus aureus. FEMS Immunol Med Microbiol.<br />

2004 Jan 15;40(1):1-9.<br />

Chida Y, Sudo N, Kubo C.<br />

Does stress exacerbate liver diseases? J Gastroenterol Hepatol. 2006 Jan;21(1 Pt 2):202-8.<br />

Chien Y, Manna AC, Projan SJ, Cheung AL.<br />

SarA, a global regulator <strong>of</strong> virulence determinants in Staphylococcus aureus, binds to a conserved motif essential for<br />

sar-dependent gene regulation. J Biol Chem. 1999 Dec 24;274(52):37169-76.<br />

Clements MO, Watson SP, Foster SJ.<br />

Characterization <strong>of</strong> the major superoxide dismutase <strong>of</strong> Staphylococcus aureus and its role in starvation survival,<br />

stress resistance, and <strong>pathogen</strong>icity. J Bacteriol. 1999 Jul;181(13):3898-903.<br />

Cohen N, Mouly E, Hamdi H, Maillot MC, Pallardy M, Godot V, Capel F, Balian A, Naveau S, Galanaud P, Lemoine FM,<br />

Emilie D.<br />

GILZ expression in human dendritic cells redirects their maturation and prevents antigen-specific T lymphocyte<br />

response. Blood. 2006 Mar 1;107(5):2037-44. Epub 2005 Nov 17.<br />

210


Maren Depke<br />

References<br />

Cole AM, Ganz T, Liese AM, Burdick MD, Liu L, Strieter RM.<br />

Cutting edge: IFN-inducible ELR – CXC chemokines display defensin-like antimicrobial activity. J Immunol. 2001 Jul 15;<br />

167(2):623-7.<br />

Cole KE, Strick CA, Paradis TJ, Ogborne KT, Loetscher M, Gladue RP, Lin W, Boyd JG, Moser B, Wood DE, Sahagan BG,<br />

Neote K.<br />

Interferon-inducible T cell alpha chemoattractant (I-TAC): a novel non-ELR CXC chemokine with potent activity on<br />

activated T cells through selective high affinity binding to CXCR3. J Exp Med. 1998 Jun 15;187(12):2009-21.<br />

Colin DA, Mazurier I, Sire S, Finck-Barbançon V.<br />

Interaction <strong>of</strong> the two components <strong>of</strong> leukocidin from Staphylococcus aureus with human polymorphonuclear<br />

leukocyte membranes: sequential binding and subsequent activation. Infect Immun. 1994 Aug;62(8):3184-8.<br />

Condon C, Squires C, Squires CL.<br />

Control <strong>of</strong> rRNA transcription in Escherichia coli. Microbiol Rev. 1995 Dec;59(4):623-45.<br />

Cook-Mills JM.<br />

VCAM-1 signals during lymphocyte migration: role <strong>of</strong> reactive oxygen species. Mol Immunol. 2002 Dec;39(9):499-<br />

508.<br />

Costelli P, Carbó N, Tessitore L, Bag<strong>by</strong> GJ, Lopez-Soriano FJ, Argilés JM, Baccino FM.<br />

Tumor necrosis factor-α mediates changes in tissue protein turnover in a rat cancer cachexia model. J Clin Invest.<br />

1993 Dec;92(6):2783-9.<br />

Crosse AM, Greenway DL, England RR.<br />

Accumulation <strong>of</strong> ppGpp and ppGp in Staphylococcus aureus 8325-4 following nutrient starvation. Lett Appl<br />

Microbiol. 2000 Oct;31(4):332-7.<br />

Cubellis MV, Wun TC, Blasi F.<br />

Receptor-mediated internalization and degradation <strong>of</strong> urokinase is caused <strong>by</strong> its specific inhibitor PAI-1. EMBO J.<br />

1990 Apr;9(4):1079-85.<br />

da Silva MC, Zahm JM, Gras D, Bajolet O, Abely M, Hinnrasky J, Milliot M, de Assis MC, Hologne C, Bonnet N,<br />

Merten M, Plotkowski MC, Puchelle E.<br />

Dynamic interaction between airway epithelial cells and Staphylococcus aureus. Am J Physiol Lung Cell Mol Physiol.<br />

2004 Sep;287(3):L543-51. Epub 2004 May 14.<br />

D'Adamio F, Zollo O, Moraca R, Ayroldi E, Bruscoli S, Bartoli A, Cannarile L, Migliorati G, Riccardi C.<br />

A new dexamethasone-induced gene <strong>of</strong> the leucine zipper family protects T lymphocytes from TCR/CD3-activated<br />

cell death. Immunity. 1997 Dec;7(6):803-12.<br />

Dallman MF.<br />

Modulation <strong>of</strong> stress responses: how we cope with excess glucocorticoids. Exp Neurol. 2007 Aug;206(2):179-82.<br />

Epub 2007 Jun 18.<br />

Dallman MF, Warne JP, Foster MT, Pecoraro NC.<br />

Glucocorticoids and insulin both modulate caloric intake through actions on the brain. J Physiol. 2007 Sep 1;<br />

583(Pt2):431-6. Epub 2007 Jun 7.<br />

Dalton DK, Pitts-Meek S, Keshav S, Figari IS, Bradley A, Stewart TA.<br />

Multiple defects <strong>of</strong> immune cell function in mice with disrupted interferon-gamma genes. Science. 1993 Mar 19;<br />

259(5102):1739-42.<br />

Dass K, Ahmad A, Azmi AS, Sarkar SH, Sarkar FH.<br />

Evolving role <strong>of</strong> uPA/uPAR system in human cancers. Cancer Treat Rev. 2008 Apr;34(2):122-36. Epub 2007 Dec 26.<br />

De Greef C, Imberechts H, Matthyssens G, Van Meirvenne N, Hamers R.<br />

A gene expressed only in serum-resistant variants <strong>of</strong> Trypanosoma brucei rhodesiense. Mol Biochem Parasitol. 1989<br />

Sep;36(2):169-76.<br />

de Haas CJ, Veldkamp KE, Peschel A, Weerkamp F, Van Wamel WJ, Heezius EC, Poppelier MJ, Van Kessel KP,<br />

van Strijp JA.<br />

Chemotaxis inhibitory protein <strong>of</strong> Staphylococcus aureus, a bacterial antiinflammatory agent. J Exp Med. 2004 Mar 1;<br />

199(5):687-95.<br />

Degrandi D, Konermann C, Beuter-Gunia C, Kresse A, Würthner J, Kurig S, Beer S, Pfeffer K.<br />

Extensive <strong>characterization</strong> <strong>of</strong> IFN-induced GTPases mGBP1 to mGBP10 involved in <strong>host</strong> defense. J Immunol. 2007<br />

Dec 1;179(11):7729-40.<br />

DeLeo FR, Allen LA, Apicella M, Nauseef WM.<br />

NADPH oxidase activation and assembly during phagocytosis. J Immunol. 1999 Dec 15;163(12):6732-40.<br />

211


Maren Depke<br />

References<br />

Demling R.<br />

The use <strong>of</strong> anabolic agents in catabolic states. J Burns Wounds. 2007 Feb 12;6:e2.<br />

Deora R, Misra TK.<br />

Purification and <strong>characterization</strong> <strong>of</strong> DNA dependent RNA polymerase from Staphylococcus aureus. Biochem Biophys<br />

Res Commun. 1995 Mar 17;208(2):610-6.<br />

Depke M, Fusch G, Domanska G, Geffers R, Völker U, Schuett C, Kiank C.<br />

Hypermetabolic syndrome as a consequence <strong>of</strong> repeated psychological stress in mice. Endocrinology. 2008 Jun;<br />

149(6):2714-23. Epub 2008 Mar 6.<br />

Depke M, Steil L, Domanska G, Völker U, Schütt C, Kiank C.<br />

Altered hepatic mRNA expression <strong>of</strong> immune response and apoptosis-associated genes after acute and chronic<br />

psychological stress in mice. Mol Immunol. 2009 Sep;46(15):3018-28. Epub 2009 Jul 9.<br />

Deurenberg RH, Stobberingh EE.<br />

The evolution <strong>of</strong> Staphylococcus aureus. Infect Genet Evol. 2008 Dec;8(6):747-63. Epub 2008 Jul 29.<br />

Diekema DJ, Pfaller MA, Schmitz FJ, Smayevsky J, Bell J, Jones RN, Beach M; SENTRY Partcipants Group.<br />

Survey <strong>of</strong> infections due to Staphylococcus species: frequency <strong>of</strong> occurrence and antimicrobial susceptibility <strong>of</strong><br />

isolates collected in the United States, Canada, Latin America, Europe, and the Western Pacific region for the<br />

SENTRY Antimicrobial Surveillance Program, 1997-1999. Clin Infect Dis. 2001 May 15;32 Suppl 2:S114-32.<br />

Dietrich CG, Geier A, Oude Elferink RP.<br />

ABC <strong>of</strong> oral bioavailability: transporters as gatekeepers in the gut. Gut. 2003 Dec;52(12):1788-95.<br />

Döffinger R, Dupuis S, Picard C, Fieschi C, Feinberg J, Barcenas-Morales G, Casanova JL.<br />

Inherited disorders <strong>of</strong> IL-12- and IFNγ-mediated immunity: a molecular genetics update. Mol Immunol. 2002 May;<br />

38(12-13):903-9.<br />

Dougall WC, Nick HS.<br />

Manganese superoxide dismutase: a hepatic acute phase protein regulated <strong>by</strong> interleukin-6 and glucocorticoids.<br />

Endocrinology. 1991 Nov;129(5):2376-84.<br />

Dubrac S, Boneca IG, Poupel O, Msadek T.<br />

New insights into the WalK/WalR (YycG/YycF) essential signal transduction pathway reveal a major role in<br />

controlling cell wall metabolism and bi<strong>of</strong>ilm formation in Staphylococcus aureus. J Bacteriol. 2007 Nov;189(22):<br />

8257-69. Epub 2007 Sep 7.<br />

Duchateau PN, Pullinger CR, Orellana RE, Kunitake ST, Naya-Vigne J, O'Connor PM, Malloy MJ, Kane JP.<br />

Apolipoprotein L, a new human high density lipoprotein apolipoprotein expressed <strong>by</strong> the pancreas. Identification,<br />

cloning, <strong>characterization</strong>, and plasma distribution <strong>of</strong> apolipoprotein L. J Biol Chem. 1997 Oct 10;272(41):25576-82.<br />

Duffield JS.<br />

The inflammatory macrophage: a story <strong>of</strong> Jekyll and Hyde. Clin Sci (Lond). 2003 Jan;104(1):27-38.<br />

Dunkelberger JR, Song WC.<br />

Complement and its role in innate and adaptive immune responses. Cell Res. 2010 Jan;20(1):34-50. Epub 2009 Dec<br />

15.<br />

Dziewanowska K, Patti JM, Deobald CF, Bayles KW, Trumble WR, Bohach GA.<br />

Fibronectin binding protein and <strong>host</strong> cell tyrosine kinase are required for internalization <strong>of</strong> Staphylococcus aureus <strong>by</strong><br />

epithelial cells. Infect Immun. 1999 Sep;67(9):4673-8.<br />

Dziewanowska K, Carson AR, Patti JM, Deobald CF, Bayles KW, Bohach GA.<br />

Staphylococcal fibronectin binding protein interacts with heat shock protein 60 and integrins: role in internalization<br />

<strong>by</strong> epithelial cells. Infect Immun. 2000 Nov;68(11):6321-8.<br />

Ehrt S, Schnappinger D, Bekiranov S, Drenkow J, Shi S, Gingeras TR, Gaasterland T, Schoolnik G, Nathan C.<br />

Reprogramming <strong>of</strong> the macrophage transcriptome in response to interferon-γ and Mycobacterium tuberculosis:<br />

signaling roles <strong>of</strong> nitric oxide synthase-2 and phagocyte oxidase. J Exp Med. 2001 Oct 15;194(8):1123-40.<br />

Elenkov IJ.<br />

Glucocorticoids and the Th1/Th2 balance. Ann N Y Acad Sci. 2004 Jun;1024:138-46.<br />

Emonts M, de Jongh CE, Houwing-Duistermaat JJ, van Leeuwen WB, de Groot R, Verbrugh HA, Hermans PW,<br />

van Belkum A.<br />

Association between nasal carriage <strong>of</strong> Staphylococcus aureus and the human complement cascade activator serine<br />

protease C1 inhibitor (C1INH) valine vs. methionine polymorphism at amino acid position 480. FEMS Immunol Med<br />

Microbiol. 2007 Aug;50(3):330-2. Epub 2007 May 10.<br />

212


Maren Depke<br />

References<br />

Eriksen NH, Espersen F, Rosdahl VT, Jensen K.<br />

Carriage <strong>of</strong> Staphylococcus aureus among 104 healthy persons during a 19-month period. Epidemiol Infect. 1995<br />

Aug;115(1):51-60.<br />

Ermak G, Davies KJ.<br />

Calcium and oxidative stress: from cell signaling to cell death. Mol Immunol. 2002 Feb;38(10):713-21.<br />

Erwig LP, Kluth DC, Walsh GM, Rees AJ.<br />

Initial cytokine exposure determines function <strong>of</strong> macrophages and renders them unresponsive to other cytokines.<br />

J Immunol. 1998 Aug 15;161(4):1983-8.<br />

Eske K, Breitbach K, Köhler J, Wongprompitak P, Steinmetz I.<br />

Generation <strong>of</strong> murine bone marrow derived macrophages in a standardised serum-free cell culture system.<br />

J Immunol Methods. 2009 Mar 15;342(1-2):13-9. Epub 2009 Jan 6.<br />

Esmon CT.<br />

Inflammation and the activated protein C anticoagulant pathway. Semin Thromb Hemost. 2006 Apr;32 Suppl 1:49-<br />

60.<br />

Eto A, Muta T, Yamazaki S, Takeshige K.<br />

Essential roles for NF-κB and a Toll/IL-1 receptor domain-specific signal(s) in the induction <strong>of</strong> IκB-ζ. Biochem Biophys<br />

Res Commun. 2003 Feb 7;301(2):495-501.<br />

Everson CA, Reed HL.<br />

Pituitary and peripheral thyroid hormone responses to thyrotropin-releasing hormone during sustained sleep<br />

deprivation in freely moving rats. Endocrinology. 1995 Apr;136(4):1426-34.<br />

Farber JM.<br />

A macrophage mRNA selectively induced <strong>by</strong> γ-interferon encodes a member <strong>of</strong> the platelet factor 4 family <strong>of</strong><br />

cytokines. Proc Natl Acad Sci U S A. 1990 Jul;87(14):5238-42.<br />

Fausto N, Campbell JS, Riehle KJ.<br />

Liver regeneration. Hepatology. 2006 Feb;43(2 Suppl 1):S45-53.<br />

Favorova OO, Zargarova TA, Rukosuyev VS, Beresten SF, Kisselev LL.<br />

Molecular and cellular studies <strong>of</strong> tryptophanyl-tRNA synthetases using monoclonal antibodies. Remarkable<br />

variations in the content <strong>of</strong> tryptophanyl-tRNA synthetase in the pancreas <strong>of</strong> different mammals. Eur J Biochem.<br />

1989 Oct 1;184(3):583-8.<br />

Feezor RJ, Oberholzer C, Baker HV, Novick D, Rubinstein M, Moldawer LL, Pribble J, Souza S, Dinarello CA, Ertel W,<br />

Oberholzer A.<br />

Molecular <strong>characterization</strong> <strong>of</strong> the acute inflammatory response to infections with Gram-negative versus Grampositive<br />

bacteria. Infect Immun. 2003 Oct;71(10):5803-13.<br />

Finck BN, Kelley KW, Dantzer R, Johnson RW.<br />

In vivo and in vitro evidence for the involvement <strong>of</strong> tumor necrosis factor-α in the induction <strong>of</strong> leptin <strong>by</strong><br />

lipopolysaccharide. Endocrinology. 1998 May;139(5):2278-83.<br />

Finck BN, Johnson RW.<br />

Tumor necrosis factor (TNF)-α induces leptin production through the p55 TNF receptor. Am J Physiol Regul Integr<br />

Comp Physiol. 2000 Feb;278(2):R537-43.<br />

Finlay BB, Cossart P.<br />

Exploitation <strong>of</strong> mammalian <strong>host</strong> cell functions <strong>by</strong> bacterial <strong>pathogen</strong>s. Science. 1997 May 2;276(5313):718-25.<br />

Fitzpatrick F, Humphreys H, O'Gara JP.<br />

The genetics <strong>of</strong> staphylococcal bi<strong>of</strong>ilm formation – will a greater understanding <strong>of</strong> <strong>pathogen</strong>esis lead to better<br />

management <strong>of</strong> device-related infection? Clin Microbiol Infect. 2005 Dec;11(12):967-73.<br />

Fleckner J, Rasmussen HH, Justesen J.<br />

Human interferon gamma potently induces the synthesis <strong>of</strong> a 55-kDa protein (gamma 2) highly homologous to<br />

rabbit peptide chain release factor and bovine tryptophanyl-tRNA synthetase. Proc Natl Acad Sci U S A. 1991 Dec 15;<br />

88(24):11520-4.<br />

Fletcher JI, Huang DC.<br />

BH3-only proteins: orchestrating cell death. Cell Death Differ. 2006 Aug;13(8):1268-71. Epub 2006 Jun 9.<br />

Fleury C, Neverova M, Collins S, Raimbault S, Champigny O, Levi-Meyrueis C, Bouillaud F, Seldin MF, Surwit RS,<br />

Ricquier D, Warden CH.<br />

Uncoupling protein-2: a novel gene linked to obesity and hyperinsulinemia. Nat Genet. 1997 Mar;15(3):269-72.<br />

213


Maren Depke<br />

References<br />

Foss GS, Prydz H.<br />

Interferon regulatory factor 1 mediates the interferon-γ induction <strong>of</strong> the human immunoproteasome subunit<br />

multicatalytic endopeptidase complex-like 1. J Biol Chem. 1999 Dec 3;274(49):35196-202.<br />

Foster TJ.<br />

Immune evasion <strong>by</strong> staphylococci. Nat Rev Microbiol. 2005 Dec;3(12):948-58.<br />

Foster TJ.<br />

Colonization and infection <strong>of</strong> the human <strong>host</strong> <strong>by</strong> staphylococci: adhesion, survival and immune evasion. Vet<br />

Dermatol. 2009 Oct;20(5-6):456-70.<br />

Foster TJ, Höök M.<br />

Surface protein adhesins <strong>of</strong> Staphylococcus aureus. Trends Microbiol. 1998 Dec;6(12):484-8.<br />

Fraser JD, Pr<strong>of</strong>t T.<br />

The bacterial superantigen and superantigen-like proteins. Immunol Rev. 2008 Oct;225:226-43.<br />

Fridovich I.<br />

The biology <strong>of</strong> oxygen radicals. Science. 1978 Sep 8;201(4359):875-80.<br />

Frolova LY, Grigorieva AY, Sudomoina MA, Kisselev LL.<br />

The human gene encoding tryptophanyl-tRNA synthetase: interferon-response elements and exon-intron<br />

organization. Gene. 1993 Jun 30;128(2):237-45.<br />

Frumento G, Rotondo R, Tonetti M, Damonte G, Benatti U, Ferrara GB.<br />

Tryptophan-derived catabolites are responsible for inhibition <strong>of</strong> T and natural killer cell proliferation induced <strong>by</strong><br />

indoleamine 2,3-dioxygenase. J Exp Med. 2002 Aug 19;196(4):459-68.<br />

Fujita T, Matsushita M, Endo Y.<br />

The lectin-complement pathway – its role in innate immunity and evolution. Immunol Rev. 2004 Apr;198:185-202.<br />

Gao S, von der Malsburg A, Paeschke S, Behlke J, Haller O, Kochs G, Daumke O.<br />

Structural basis <strong>of</strong> oligomerization in the stalk region <strong>of</strong> dynamin-like MxA. Nature. 2010 May 27;465(7297):502-6.<br />

Epub 2010 Apr 28.<br />

Garg AK, Aggarwal BB.<br />

Reactive oxygen intermediates in TNF signaling. Mol Immunol. 2002 Dec;39(9):509-17.<br />

Garzoni C, Francois P, Huyghe A, Couzinet S, Tapparel C, Charbonnier Y, Renzoni A, Lucchini S, Lew DP, Vaudaux P,<br />

Kelley WL, Schrenzel J.<br />

A global view <strong>of</strong> Staphylococcus aureus whole genome expression upon internalization in human epithelial cells.<br />

BMC Genomics. 2007 Jun 14;8:171.<br />

Gaur U, Aggarwal BB.<br />

Regulation <strong>of</strong> proliferation, survival and apoptosis <strong>by</strong> members <strong>of</strong> the TNF superfamily. Biochem Pharmacol. 2003<br />

Oct 15;66(8):1403-8.<br />

Ghosh S, Preet A, Groopman JE, Ganju RK.<br />

Cannabinoid receptor CB 2 modulates the CXCL12/CXCR4-mediated chemotaxis <strong>of</strong> T lymphocytes. Mol Immunol.<br />

2006 Jul;43(14):2169-79. Epub 2006 Feb 28.<br />

Giachino P, Engelmann S, Bisch<strong>of</strong>f M.<br />

σ B activity depends on RsbU in Staphylococcus aureus. J Bacteriol. 2001 Mar;183(6):1843-52.<br />

Gibbs DF, Warner RL, Weiss SJ, Johnson KJ, Varani J.<br />

Characterization <strong>of</strong> matrix metalloproteinases produced <strong>by</strong> rat alveolar macrophages. Am J Respir Cell Mol Biol.<br />

1999 Jun;20(6):1136-44.<br />

Gibbs DF, Shanley TP, Warner RL, Murphy HS, Varani J, Johnson KJ.<br />

Role <strong>of</strong> matrix metalloproteinases in models <strong>of</strong> macrophage-dependent acute lung injury. Evidence for alveolar<br />

macrophage as source <strong>of</strong> proteinases. Am J Respir Cell Mol Biol. 1999 Jun;20(6):1145-54.<br />

Gill SR, Fouts DE, Archer GL, Mongodin EF, Deboy RT, Ravel J, Paulsen IT, Kolonay JF, Brinkac L, Beanan M, Dodson RJ,<br />

Daugherty SC, Madupu R, Angiuoli SV, Durkin AS, Haft DH, Vamathevan J, Khouri H, Utterback T, Lee C, Dimitrov G,<br />

Jiang L, Qin H, Weidman J, Tran K, Kang K, Hance IR, Nelson KE, Fraser CM.<br />

Insights on evolution <strong>of</strong> virulence and resistance from the complete genome analysis <strong>of</strong> an early methicillin-resistant<br />

Staphylococcus aureus strain and a bi<strong>of</strong>ilm-producing methicillin-resistant Staphylococcus epidermidis strain.<br />

J Bacteriol. 2005 Apr;187(7):2426-38.<br />

214


Maren Depke<br />

References<br />

Glaser R, Friedman SB, Smyth J, Ader R, Bijur P, Brunell P, Cohen N, Krilov LR, Lifrak ST, Stone A, T<strong>of</strong>fler P.<br />

The differential impact <strong>of</strong> training stress and final examination stress on herpesvirus latency at the United States<br />

Military Academy at West Point. Brain Behav Immun. 1999 Sep;13(3):240-51.<br />

Gleeson M, McDonald WA, Cripps AW, Pyne DB, Clancy RL, Fricker PA.<br />

The effect on immunity <strong>of</strong> long-term intensive training in elite swimmers. Clin Exp Immunol. 1995 Oct;102(1):210-6.<br />

Gobin SJ, van den Elsen PJ.<br />

Transcriptional regulation <strong>of</strong> the MHC class Ib genes HLA-E, HLA-F, and HLA-G. Hum Immunol. 2000 Nov;<br />

61(11):1102-7.<br />

Goerke C, Fluckiger U, Steinhuber A, Bisanzio V, Ulrich M, Bisch<strong>of</strong>f M, Patti JM, Wolz C.<br />

Role <strong>of</strong> Staphylococcus aureus global regulators sae and σ B in virulence gene expression during device-related<br />

infection. Infect Immun. 2005 Jun;73(6):3415-21.<br />

Goodman MN.<br />

Tumor necrosis factor induces skeletal muscle protein breakdown in rats. Am J Physiol. 1991 May;260(5 Pt 1):<br />

E727-30.<br />

Gordon RJ, Lowy FD.<br />

Pathogenesis <strong>of</strong> methicillin-resistant Staphylococcus aureus infection. Clin Infect Dis. 2008 Jun 1;46 Suppl 5:S350-9.<br />

Gordon S.<br />

The macrophage: past, present and future. Eur J Immunol. 2007 Nov;37 Suppl 1:S9-17.<br />

Goto H, Shimada K, Ikemoto H, Oguri T; Study Group on Antimicrobial Susceptibility <strong>of</strong> Pathogens Isolated from<br />

Respiratory Infections.<br />

Antimicrobial susceptibility <strong>of</strong> <strong>pathogen</strong>s isolated from more than 10,000 patients with infectious respiratory<br />

diseases: a 25-year longitudinal study. J Infect Chemother. 2009 Dec;15(6):347-60.<br />

Gracie JA, Robertson SE, McInnes IB.<br />

Interleukin-18. J Leukoc Biol. 2003 Feb;73(2):213-24.<br />

Gravet A, Colin DA, Keller D, Girardot R, Monteil H, Prévost G.<br />

Characterization <strong>of</strong> a novel structural member, LukE-LukD, <strong>of</strong> the bi-component staphylococcal leucotoxins family.<br />

FEBS Lett. 1998 Oct 2;436(2):202-8.<br />

Greenberg S, Grinstein S.<br />

Phagocytosis and innate immunity. Curr Opin Immunol. 2002 Feb;14(1):136-45.<br />

Grinholc M, Wegrzyn G, Kurlenda J.<br />

Evaluation <strong>of</strong> bi<strong>of</strong>ilm production and prevalence <strong>of</strong> the icaD gene in methicillin-resistant and methicillin-susceptible<br />

Staphylococcus aureus strains isolated from patients with nosocomial infections and carriers. FEMS Immunol Med<br />

Microbiol. 2007 Aug;50(3):375-9. Epub 2007 May 30.<br />

Gross M, Cramton SE, Götz F, Peschel A.<br />

Key role <strong>of</strong> teichoic acid net charge in Staphylococcus aureus colonization <strong>of</strong> artificial surfaces. Infect Immun. 2001<br />

May;69(5):3423-6.<br />

Grundmeier M, Tuchscherr L, Brück M, Viemann D, Roth J, Willscher E, Becker K, Peters G, Löffler B.<br />

Staphylococcal strains vary greatly in their ability to induce an inflammatory response in endothelial cells. J Infect<br />

Dis. 2010 Mar 15;201(6):871-80.<br />

Haddad JJ, Harb HL.<br />

L-γ-Glutamyl-L-cysteinyl-glycine (glutathione; GSH) and GSH-related enzymes in the regulation <strong>of</strong> pro- and antiinflammatory<br />

cytokines: a signaling transcriptional scenario for redox(y) immunologic sensor(s)? Mol Immunol. 2005<br />

May;42(9):987-1014. Epub 2004 Nov 23.<br />

Haggar A, Ehrnfelt C, Holgersson J, Flock JI.<br />

The extracellular adherence protein from Staphylococcus aureus inhibits neutrophil binding to endothelial cells.<br />

Infect Immun. 2004 Oct;72(10):6164-7.<br />

Haggar A, Hussain M, Lönnies H, Herrmann M, Norr<strong>by</strong>-Teglund A, Flock JI.<br />

Extracellular adherence protein from Staphylococcus aureus enhances internalization into eukaryotic cells. Infect<br />

Immun. 2003 May;71(5):2310-7.<br />

Haggar A, Shannon O, Norr<strong>by</strong>-Teglund A, Flock JI.<br />

Dual effects <strong>of</strong> extracellular adherence protein from Staphylococcus aureus on peripheral blood mononuclear cells.<br />

J Infect Dis. 2005 Jul 15;192(2):210-7. Epub 2005 Jun 13.<br />

215


Maren Depke<br />

References<br />

Hagopian K, Ramsey JJ, Weindruch R.<br />

Caloric restriction increases gluconeogenic and transaminase enzyme activities in mouse liver. Exp Gerontol. 2003<br />

Mar;38(3):267-78.<br />

Haller O, Kochs G.<br />

Interferon-induced Mx proteins: dynamin-like GTPases with antiviral activity. Traffic. 2002 Oct;3(10):710-7.<br />

Haller O, Staeheli P, Kochs G.<br />

Interferon-induced Mx proteins in antiviral <strong>host</strong> defense. Biochimie. 2007 Jun-Jul;89(6-7):812-8. Epub 2007 May 8.<br />

Hallermalm K, Seki K, Wei C, Castelli C, Rivoltini L, Kiessling R, Levitskaya J.<br />

Tumor necrosis factor-α induces coordinated changes in major histocompatibility class I presentation pathway,<br />

resulting in increased stability <strong>of</strong> class I complexes at the cell surface. Blood. 2001 Aug 15;98(4):1108-15.<br />

Hammel M, Sfyroera G, Pyrpassopoulos S, Ricklin D, Ramyar KX, Pop M, Jin Z, Lambris JD, Geisbrecht BV.<br />

Characterization <strong>of</strong> Ehp, a secreted complement inhibitory protein from Staphylococcus aureus. J Biol Chem. 2007<br />

Oct 12;282(41):30051-61. Epub 2007 Aug 15.<br />

Hang CH, Shi JX, Li JS, Wu W, Yin HX.<br />

Alterations <strong>of</strong> intestinal mucosa structure and barrier function following traumatic brain injury in rats. World J<br />

Gastroenterol. 2003 Dec;9(12):2776-81.<br />

Hansen MB, Dresner LS, Wait RB.<br />

Pr<strong>of</strong>ile <strong>of</strong> neurohumoral agents on mesenteric and intestinal blood flow in health and disease. Physiol Res. 1998;<br />

47(5):307-27.<br />

Hansen TH, Huang S, Arnold PL, Fremont DH.<br />

Patterns <strong>of</strong> nonclassical MHC antigen presentation. Nat Immunol. 2007 Jun;8(6):563-8.<br />

Hara T, Ogasawara N, Akimoto H, Takikawa O, Hiramatsu R, Kawabe T, Isobe K, Nagase F.<br />

High-affinity uptake <strong>of</strong> kynurenine and nitric oxide-mediated inhibition <strong>of</strong> indoleamine 2,3-dioxygenase in bone<br />

marrow-derived myeloid dendritic cells. Immunol Lett. 2008 Feb 15;116(1):95-102. Epub 2007 Dec 26.<br />

Harraghy N, Hussain M, Haggar A, Chavakis T, Sinha B, Herrmann M, Flock JI.<br />

The adhesive and immunomodulating properties <strong>of</strong> the multifunctional Staphylococcus aureus protein Eap.<br />

Microbiology. 2003 Oct;149(Pt 10):2701-7.<br />

Harraghy N, Kormanec J, Wolz C, Homerova D, Goerke C, Ohlsen K, Qazi S, Hill P, Herrmann M.<br />

sae is essential for expression <strong>of</strong> the staphylococcal adhesins Eap and Emp. Microbiology. 2005 Jun;151(Pt 6):1789-<br />

800.<br />

Harris RB, Zhou J, Youngblood BD, Rybkin II, Smagin GN, Ryan DH.<br />

Effect <strong>of</strong> repeated stress on body weight and body composition <strong>of</strong> rats fed low- and high-fat diets. Am J Physiol.<br />

1998 Dec;275(6 Pt 2):R1928-38.<br />

Harris RB, Mitchell TD, Simpson J, Redmann SM Jr, Youngblood BD, Ryan DH.<br />

Weight loss in rats exposed to repeated acute restraint stress is independent <strong>of</strong> energy or leptin status. Am J Physiol<br />

Regul Integr Comp Physiol. 2002 Jan;282(1):R77-88.<br />

Harris RB, Palmondon J, Leshin S, Flatt WP, Richard D.<br />

Chronic disruption <strong>of</strong> body weight but not <strong>of</strong> stress peptides or receptors in rats exposed to repeated restraint<br />

stress. Horm Behav. 2006 May;49(5):615-25. Epub 2006 Jan 19.<br />

Hartleib J, Köhler N, Dickinson RB, Chhatwal GS, Sixma JJ, Hartford OM, Foster TJ, Peters G, Kehrel BE, Herrmann M.<br />

Protein A is the von Willebrand factor binding protein on Staphylococcus aureus. Blood. 2000 Sep 15;96(6):2149-56.<br />

Hauck CR, Ohlsen K.<br />

Sticky connections: extracellular matrix protein recognition and integrin-mediated cellular invasion <strong>by</strong><br />

Staphylococcus aureus. Curr Opin Microbiol. 2006 Feb;9(1):5-11. Epub 2006 Jan 6.<br />

Hehlgans T, Pfeffer K.<br />

The intriguing biology <strong>of</strong> the tumour necrosis factor/tumour necrosis factor receptor superfamily: players, rules and<br />

the games. Immunology. 2005 May;115(1):1-20.<br />

Helbig KJ, Ruszkiewicz A, Semendric L, Harley HA, McColl SR, Beard MR.<br />

Expression <strong>of</strong> the CXCR3 ligand I-TAC <strong>by</strong> hepatocytes in chronic hepatitis C and its correlation with hepatic<br />

inflammation. Hepatology. 2004 May;39(5):1220-9.<br />

Herbert S, Ziebandt AK, Ohlsen K, Schäfer T, Hecker M, Albrecht D, Novick R, Götz F.<br />

Repair <strong>of</strong> global regulators in Staphylococcus aureus 8325 and comparative analysis with other clinical isolates.<br />

Infect Immun. 2010 Jun;78(6):2877-89. Epub 2010 Mar 8.<br />

216


Maren Depke<br />

References<br />

Herman A, Kappler JW, Marrack P, Pullen AM.<br />

Superantigens: mechanism <strong>of</strong> T-cell stimulation and role in immune responses. Annu Rev Immunol. 1991;9:745-72.<br />

Herold MJ, McPherson KG, Reichardt HM.<br />

Glucocorticoids in T cell apoptosis and function. Cell Mol Life Sci. 2006 Jan;63(1):60-72.<br />

Hillgartner FB, Salati LM, Goodridge AG.<br />

Physiological and molecular mechanisms involved in nutritional regulation <strong>of</strong> fatty acid synthesis. Physiol Rev. 1995<br />

Jan;75(1):47-76.<br />

Hiramatsu K, Watanabe S, Takeuchi F, Ito T, Baba T.<br />

Genetic <strong>characterization</strong> <strong>of</strong> methicillin-resistant Staphylococcus aureus. Vaccine. 2004 Dec 6;22 Suppl 1:S5-8.<br />

Horsburgh MJ, Wharton SJ, Cox AG, Ingham E, Peacock S, Foster SJ.<br />

MntR modulates expression <strong>of</strong> the PerR regulon and superoxide resistance in Staphylococcus aureus through<br />

control <strong>of</strong> manganese uptake. Mol Microbiol. 2002 Jun;44(5):1269-86.<br />

Horsburgh MJ, Aish JL, White IJ, Shaw L, Lithgow JK, Foster SJ.<br />

σ B modulates virulence determinant expression and stress resistance: <strong>characterization</strong> <strong>of</strong> a functional rsbU strain<br />

derived from Staphylococcus aureus 8325-4. J Bacteriol. 2002 Oct;184(19):5457-67.<br />

Hu X, Chakravarty SD, Ivashkiv LB.<br />

Regulation <strong>of</strong> interferon and Toll-like receptor signaling during macrophage activation <strong>by</strong> opposing feedforward and<br />

feedback inhibition mechanisms. Immunol Rev. 2008 Dec;226:41-56.<br />

Huang S, Hendriks W, Althage A, Hemmi S, Bluethmann H, Kamijo R, Vilcek J, Zinkernagel RM, Aguet M.<br />

Immune response in mice that lack the interferon-gamma receptor. Science. 1993 Mar 19;259(5102):1742-5.<br />

Hucke C, MacKenzie CR, Adjogble KD, Takikawa O, Däubener W.<br />

Nitric oxide-mediated regulation <strong>of</strong> gamma interferon-induced bacteriostasis: inhibition and degradation <strong>of</strong> human<br />

indoleamine 2,3-dioxygenase. Infect Immun. 2004 May;72(5):2723-30.<br />

Hudson MC, Ramp WK, Nicholson NC, Williams AS, Nousiainen MT.<br />

Internalization <strong>of</strong> Staphylococcus aureus <strong>by</strong> cultured osteoblasts. Microb Pathog. 1995 Dec;19(6):409-19.<br />

Hultgren O, Eugster HP, Sedgwick JD, Körner H, Tarkowski A.<br />

TNF/lymphotoxin-α double-mutant mice resist septic arthritis but display increased mortality in response to<br />

Staphylococcus aureus. J Immunol. 1998 Dec 1;161(11):5937-42.<br />

Hussain M, Haggar A, Heilmann C, Peters G, Flock JI, Herrmann M.<br />

Insertional inactivation <strong>of</strong> eap in Staphylococcus aureus strain Newman confers reduced staphylococcal binding to<br />

fibroblasts. Infect Immun. 2002 Jun;70(6):2933-40.<br />

Hussain M, Schäfer D, Juuti KM, Peters G, Haslinger-Löffler B, Kuusela PI, Sinha B.<br />

Expression <strong>of</strong> Pls (plasmin sensitive) in Staphylococcus aureus negative for pls reduces adherence and cellular<br />

invasion and acts <strong>by</strong> steric hindrance. J Infect Dis. 2009 Jul 1;200(1):107-17.<br />

Iandolo JJ, Worrell V, Groicher KH, Qian Y, Tian R, Kenton S, Dorman A, Ji H, Lin S, Loh P, Qi S, Zhu H, Roe BA.<br />

Comparative analysis <strong>of</strong> the genomes <strong>of</strong> the temperate bacteriophages phi 11, phi 12 and phi 13 <strong>of</strong> Staphylococcus<br />

aureus 8325. Gene. 2002 May 1;289(1-2):109-18.<br />

Iber H, Li-Masters T, Chen Q, Yu S, Morgan ET.<br />

Regulation <strong>of</strong> hepatic cytochrome P450 2C11 via cAMP: implications for down-regulation in diabetes, fasting, and<br />

inflammation. J Pharmacol Exp Ther. 2001 Apr;297(1):174-80.<br />

Iles KE, Forman HJ.<br />

Macrophage signaling and respiratory burst. Immunol Res. 2002;26(1-3):95-105.<br />

Imlay JA, Linn S.<br />

DNA damage and oxygen radical toxicity. Science. 1988 Jun 3;240(4857):1302-9.<br />

Janeway CA Jr, Medzhitov R.<br />

Innate immune recognition. Annu Rev Immunol. 2002;20:197-216. Epub 2001 Oct 4.<br />

Janeway CA Jr., Travers P, Walport M, Shlomchik MJ.<br />

Immunologie. 5. Auflage, 2002, Spektrum Akademischer Verlag GmbH Heidelberg Berlin. ISBN 3-8274-1079-7.<br />

Jensen PE.<br />

Mechanisms <strong>of</strong> antigen presentation. Clin Chem Lab Med. 1999 Mar;37(3):179-86.<br />

217


Maren Depke<br />

References<br />

Jin T, Bokarewa M, Foster T, Mitchell J, Higgins J, Tarkowski A.<br />

Staphylococcus aureus resists human defensins <strong>by</strong> production <strong>of</strong> staphylokinase, a novel bacterial evasion<br />

mechanism. J Immunol. 2004 Jan 15;172(2):1169-76.<br />

Johnson LM, Scott P.<br />

STAT1 expression in dendritic cells, but not T cells, is required for immunity to Leishmania major. J Immunol. 2007<br />

Jun 1;178(11):7259-66.<br />

Jonsson IM, Arvidson S, Foster S, Tarkowski A.<br />

Sigma factor B and RsbU are required for virulence in Staphylococcus aureus-induced arthritis and sepsis. Infect<br />

Immun. 2004 Oct;72(10):6106-11.<br />

Jönsson K, McDevitt D, McGavin MH, Patti JM, Höök M.<br />

Staphylococcus aureus expresses a major histocompatibility complex class II analog. J Biol Chem. 1995 Sep 15;<br />

270(37):21457-60.<br />

Jung T, Catalgol B, Grune T.<br />

The proteasomal system. Mol Aspects Med. 2009 Aug;30(4):191-296. Epub 2009 Apr 14.<br />

Jurado J, Prieto-Alamo MJ, Madrid-Rísquez J, Pueyo C.<br />

Absolute gene expression patterns <strong>of</strong> thioredoxin and glutaredoxin redox systems in mouse. J Biol Chem. 2003 Nov<br />

14;278(46):45546-54. Epub 2003 Sep 3.<br />

Kaliman PA, Barannik T, Strel'chenko E, Inshina N, Sokol O.<br />

Intracellular redistribution <strong>of</strong> heme in rat liver under oxidative stress: the role <strong>of</strong> heme synthesis. Cell Biol Int. 2005<br />

Jan;29(1):9-14. Epub 2005 Jan 26.<br />

Kaneko J, Kamio Y.<br />

Bacterial two-component and hetero-heptameric pore-forming cytolytic toxins: structures, pore-forming<br />

mechanism, and organization <strong>of</strong> the genes. Biosci Biotechnol Biochem. 2004 May;68(5):981-1003.<br />

Kantari C, Pederzoli-Ribeil M, Witko-Sarsat V.<br />

The role <strong>of</strong> neutrophils and monocytes in innate immunity. Contrib Microbiol. 2008;15:118-46.<br />

Karavolos MH, Horsburgh MJ, Ingham E, Foster SJ.<br />

Role and regulation <strong>of</strong> the superoxide dismutases <strong>of</strong> Staphylococcus aureus. Microbiology. 2003 Oct;149(Pt 10):<br />

2749-58.<br />

Kawabata S, Iwanaga S.<br />

Structure and function <strong>of</strong> staphylothrombin. Semin Thromb Hemost. 1994;20(4):345-50.<br />

Kawabata S, Morita T, Iwanaga S, Igarashi H.<br />

Enzymatic properties <strong>of</strong> staphylothrombin, an active molecular complex formed between staphylocoagulase and<br />

human prothrombin. J Biochem. 1985 Dec;98(6):1603-14.<br />

Khan S, van den Broek M, Schwarz K, de Giuli R, Diener PA, Groettrup M.<br />

Immunoproteasomes largely replace constitutive proteasomes during an antiviral and antibacterial immune<br />

response in the liver. J Immunol. 2001 Dec 15;167(12):6859-68.<br />

Khan WA, Blobe GC, Hannun YA.<br />

Arachidonic acid and free fatty acids as second messengers and the role <strong>of</strong> protein kinase C. Cell Signal. 1995<br />

Mar;7(3):171-84.<br />

Kiank C, Holtfreter B, Starke A, Mundt A, Wilke C, Schütt C.<br />

Stress susceptibility predicts the severity <strong>of</strong> immune depression and the failure to combat bacterial infections in<br />

chronically stressed mice. Brain Behav Immun. 2006 Jul;20(4):359-68. Epub 2005 Dec 2.<br />

Kiank C, Entleutner M, Fürll B, Westerholt A, Heidecke CD, Schütt C.<br />

Stress-induced immune conditioning affects the course <strong>of</strong> experimental peritonitis. Shock. 2007 Mar;27(3):305-11.<br />

Kiank C, Koerner P, Kessler W, Traeger T, Maier S, Heidecke CD, Schuett C.<br />

Seasonal variations in inflammatory responses to sepsis and stress in mice. Crit Care Med. 2007 Oct;35(10):2352-8.<br />

Kiank C, Daeschlein G, Schuett C.<br />

Pneumonia as a long-term consequence <strong>of</strong> chronic psychological stress in BALB/c mice. Brain Behav Immun. 2008<br />

Nov;22(8):1173-7. Epub 2008 Jun 20.<br />

Kiank C, Zeden JP, Drude S, Domanska G, Fusch G, Otten W, Schuett C.<br />

Psychological stress-induced, IDO1-dependent tryptophan catabolism: implications on immunosuppression in mice<br />

and humans. PLoS One. 2010 Jul 28;5(7):e11825.<br />

218


Maren Depke<br />

References<br />

Kim JS, Kim JG, Moon MY, Jeon CY, Won HY, Kim HJ, Jeon YJ, Seo JY, Kim JI, Kim J, Lee JY, Kim PH, Park JB.<br />

Transforming growth factor-β1 regulates macrophage migration via RhoA. Blood. 2006 Sep 15;108(6):1821-9. Epub<br />

2006 May 16.<br />

King NJ, Thomas SR.<br />

Molecules in focus: indoleamine 2,3-dioxygenase. Int J Biochem Cell Biol. 2007;39(12):2167-72. Epub 2007 Jan 20.<br />

Kisselev LL.<br />

Mammalian tryptophanyl-tRNA synthetases. Biochimie. 1993;75(12):1027-39.<br />

Klamp T, Boehm U, Schenk D, Pfeffer K, Howard JC.<br />

A giant GTPase, very large inducible GTPase-1, is inducible <strong>by</strong> IFNs. J Immunol. 2003 Aug 1;171(3):1255-65.<br />

Klein C, Wüstefeld T, Assmus U, Roskams T, Rose-John S, Müller M, Manns MP, Ernst M, Trautwein C.<br />

The IL-6-gp130-STAT3 pathway in hepatocytes triggers liver protection in T cell-mediated liver injury. J Clin Invest.<br />

2005 Apr;115(4):860-9. Epub 2005 Mar 3.<br />

Kluytmans J, van Belkum A, Verbrugh H.<br />

Nasal carriage <strong>of</strong> Staphylococcus aureus: epidemiology, underlying mechanisms, and associated risks. Clin Microbiol<br />

Rev. 1997 Jul;10(3):505-20. Review.<br />

Knowles RG, Moncada S.<br />

Nitric oxide synthases in mammals. Biochem J. 1994 Mar 1;298 ( Pt 2):249-58.<br />

Koban M, Swinson KL.<br />

Chronic REM-sleep deprivation <strong>of</strong> rats elevates metabolic rate and increases UCP1 gene expression in brown<br />

adipose tissue. Am J Physiol Endocrinol Metab. 2005 Jul;289(1):E68-74. Epub 2005 Feb 22.<br />

Kohler C, Wolff S, Albrecht D, Fuchs S, Becher D, Büttner K, Engelmann S, Hecker M.<br />

Proteome analyses <strong>of</strong> Staphylococcus aureus in growing and non-growing cells: a physiological approach. Int J Med<br />

Microbiol. 2005 Dec;295(8):547-65. Epub 2005 Oct 25.<br />

Kokai-Kun JF, Walsh SM, Chanturiya T, Mond JJ.<br />

Lysostaphin cream eradicates Staphylococcus aureus nasal colonization in a cotton rat model. Antimicrob Agents<br />

Chemother. 2003 May;47(5):1589-97.<br />

König B, Prévost G, König W.<br />

Composition <strong>of</strong> staphylococcal bi-component toxins determines pathophysiological reactions. J Med Microbiol. 1997<br />

Jun;46(6):479-85.<br />

Koolhaas JM, de Boer SF, Buwalda B, van Reenen K.<br />

Individual variation in coping with stress: a multidimensional approach <strong>of</strong> ultimate and proximate mechanisms.<br />

Brain Behav Evol. 2007;70(4):218-26. Epub 2007 Sep 18.<br />

Kota RS, Rutledge JC, Gohil K, Kumar A, Enelow RI, Ramana CV.<br />

Regulation <strong>of</strong> gene expression in RAW 264.7 macrophage cell line <strong>by</strong> interferon-γ. Biochem Biophys Res Commun.<br />

2006 Apr 21;342(4):1137-46. Epub 2006 Feb 24.<br />

Kullik I, Giachino P.<br />

The alternative sigma factor σ B in Staphylococcus aureus: regulation <strong>of</strong> the sigB operon in response to growth phase<br />

and heat shock. Arch Microbiol. 1997 Mar 7;167(2/3):151-9.<br />

Kullik I, Giachino P, Fuchs T.<br />

Deletion <strong>of</strong> the alternative sigma factor σ B in Staphylococcus aureus reveals its function as a global regulator <strong>of</strong><br />

virulence genes. J Bacteriol. 1998 Sep;180(18):4814-20.<br />

Kuo LE, Abe K, Zukowska Z.<br />

Stress, NPY and vascular remodeling: Implications for stress-related diseases. Peptides. 2007 Feb;28(2):435-40. Epub<br />

2007 Jan 22.<br />

Kuroda M, Kuwahara-Arai K, Hiramatsu K.<br />

Identification <strong>of</strong> the up- and down-regulated genes in vancomycin-resistant Staphylococcus aureus strains Mu3 and<br />

Mu50 <strong>by</strong> cDNA differential hybridization method. Biochem Biophys Res Commun. 2000 Mar 16;269(2):485-90.<br />

Kuroda M, Kuroda H, Oshima T, Takeuchi F, Mori H, Hiramatsu K.<br />

Two-component system VraSR positively modulates the regulation <strong>of</strong> cell-wall biosynthesis pathway in<br />

Staphylococcus aureus. Mol Microbiol. 2003 Aug;49(3):807-21.<br />

Kuypers JM, Proctor RA.<br />

Reduced adherence to traumatized rat heart valves <strong>by</strong> a low-fibronectin-binding mutant <strong>of</strong> Staphylococcus aureus.<br />

Infect Immun. 1989 Aug;57(8):2306-12.<br />

219


Maren Depke<br />

References<br />

Kwan T, Liu J, DuBow M, Gros P, Pelletier J.<br />

The complete genomes and proteomes <strong>of</strong> 27 Staphylococcus aureus bacteriophages. Proc Natl Acad Sci U S A. 2005<br />

Apr 5;102(14):5174-9. Epub 2005 Mar 23.<br />

Lam TK, Gutierrez-Juarez R, Pocai A, Bhanot S, Tso P, Schwartz GJ, Rossetti L.<br />

Brain glucose metabolism controls the hepatic secretion <strong>of</strong> triglyceride-rich lipoproteins. Nat Med. 2007 Feb;<br />

13(2):171-80. Epub 2007 Feb 4.<br />

Lambert GW, Straznicky NE, Lambert EA, Dixon JB, Schlaich MP.<br />

Sympathetic nervous activation in obesity and the metabolic syndrome – Causes, consequences and therapeutic<br />

implications. Pharmacol Ther. 2010 May;126(2):159-72. Epub 2010 Feb 19.<br />

Lang CH, Bag<strong>by</strong> GJ, Spitzer JJ.<br />

Carbohydrate dynamics in the hypermetabolic septic rat. Metabolism. 1984 Oct;33(10):959-63.<br />

Laye JP, Gill JH.<br />

Phospholipase A2 expression in tumours: a target for therapeutic intervention? Drug Discov Today. 2003 Aug1;<br />

8(15):710-6.<br />

Le Bail O, Schmidt-Ullrich R, Israël A.<br />

Promoter analysis <strong>of</strong> the gene encoding the IκB-α/MAD3 inhibitor <strong>of</strong> NF-κB: positive regulation <strong>by</strong> members <strong>of</strong> the<br />

rel/NF-κB family. EMBO J. 1993 Dec 15;12(13):5043-9.<br />

Lecordier L, Vanhollebeke B, Poelvoorde P, Tebabi P, Paturiaux-Hanocq F, Andris F, Lins L, Pays E.<br />

C-terminal mutants <strong>of</strong> apolipoprotein L-I efficiently kill both Trypanosoma brucei brucei and Trypanosoma brucei<br />

rhodesiense. PLoS Pathog. 2009 Dec;5(12):e1000685. Epub 2009 Dec 4.<br />

Lee LY, Miyamoto YJ, McIntyre BW, Höök M, McCrea KW, McDevitt D, Brown EL.<br />

The Staphylococcus aureus Map protein is an immunomodulator that interferes with T cell-mediated responses.<br />

J Clin Invest. 2002 Nov;110(10):1461-71.<br />

Lee LY, Höök M, Haviland D, Wetsel RA, Yonter EO, Syribeys P, Vernachio J, Brown EL.<br />

Inhibition <strong>of</strong> complement activation <strong>by</strong> a secreted Staphylococcus aureus protein. J Infect Dis. 2004 Aug 1;<br />

190(3):571-9. Epub 2004 Jul 6.<br />

Lee LY, Liang X, Höök M, Brown EL.<br />

Identification and <strong>characterization</strong> <strong>of</strong> the C3 binding domain <strong>of</strong> the Staphylococcus aureus extracellular fibrinogenbinding<br />

protein (Efb). J Biol Chem. 2004 Dec 3;279(49):50710-6. Epub 2004 Aug 26.<br />

Leibowitz SF, Wortley KE.<br />

Hypothalamic control <strong>of</strong> energy balance: different peptides, different functions. Peptides. 2004 Mar;25(3):473-504.<br />

Leon LR.<br />

Invited review: cytokine regulation <strong>of</strong> fever: studies using gene knockout mice. J Appl Physiol. 2002 Jun;92(6):2648-<br />

55.<br />

Leonard BE.<br />

HPA and immune axes in stress: involvement <strong>of</strong> the serotonergic system. Neuroimmunomodulation. 2006;13(5-6):<br />

268-76. Epub 2007 Aug 6.<br />

Levels JH, Lemaire LC, van den Ende AE, van Deventer SJ, van Lanschot JJ.<br />

Lipid composition and lipopolysaccharide binding capacity <strong>of</strong> lipoproteins in plasma and lymph <strong>of</strong> patients with<br />

systemic inflammatory response syndrome and multiple organ failure. Crit Care Med. 2003 Jun;31(6):1647-53.<br />

Li G, Zhang J, Sun Y, Wang H, Wang Y.<br />

The evolutionarily dynamic IFN-inducible GTPase proteins play conserved immune functions in vertebrates and<br />

cephalochordates. Mol Biol Evol. 2009 Jul;26(7):1619-30. Epub 2009 Apr 15.<br />

Liang X, Yu C, Sun J, Liu H, Landwehr C, Holmes D, Ji Y.<br />

Inactivation <strong>of</strong> a two-component signal transduction system, SaeRS, eliminates adherence and attenuates virulence<br />

<strong>of</strong> Staphylococcus aureus. Infect Immun. 2006 Aug;74(8):4655-65.<br />

Liao F, Rabin RL, Yannelli JR, Koniaris LG, Vanguri P, Farber JM.<br />

Human Mig chemokine: biochemical and functional <strong>characterization</strong>. J Exp Med. 1995 Nov 1;182(5):1301-14.<br />

Li-Hawkins J, Lund EG, Bronson AD, Russell DW.<br />

Expression cloning <strong>of</strong> an oxysterol 7α-hydroxylase selective for 24-hydroxycholesterol. J Biol Chem. 2000 Jun 2;<br />

275(22):16543-9.<br />

220


Maren Depke<br />

References<br />

Limmer A, Ohl J, Kurts C, Ljunggren HG, Reiss Y, Groettrup M, Momburg F, Arnold B, Knolle PA.<br />

Efficient presentation <strong>of</strong> exogenous antigen <strong>by</strong> liver endothelial cells to CD8+ T cells results in antigen-specific T-cell<br />

tolerance. Nat Med. 2000 Dec;6(12):1348-54.<br />

Lin CH, Sheu SY, Lee HM, Ho YS, Lee WS, Ko WC, Sheu JR.<br />

Involvement <strong>of</strong> protein kinase C-γ in IL-1β-induced cyclooxygenase-2 expression in human pulmonary epithelial<br />

cells. Mol Pharmacol. 2000 Jan;57(1):36-43.<br />

Lin CH, Kuan IH, Lee HM, Lee WS, Sheu JR, Ho YS, Wang CH, Kuo HP.<br />

Induction <strong>of</strong> cyclooxygenase-2 protein <strong>by</strong> lipoteichoic acid from Staphylococcus aureus in human pulmonary<br />

epithelial cells: involvement <strong>of</strong> a nuclear factor-κB-dependent pathway. Br J Pharmacol. 2001 Oct;134(3):543-52.<br />

Lindsay JA, Holden MT.<br />

Staphylococcus aureus: superbug, super genome? Trends Microbiol. 2004 Aug;12(8):378-85.<br />

Lindsay JA, Ruzin A, Ross HF, Kurepina N, Novick RP.<br />

The gene for toxic shock toxin is carried <strong>by</strong> a family <strong>of</strong> mobile <strong>pathogen</strong>icity islands in Staphylococcus aureus. Mol<br />

Microbiol. 1998 Jul;29(2):527-43.<br />

Linge HM, Collin M, Giwercman A, Malm J, Bjartell A, Egesten A.<br />

The antibacterial chemokine MIG/CXCL9 is constitutively expressed in epithelial cells <strong>of</strong> the male urogenital tract<br />

and is present in seminal plasma. J Interferon Cytokine Res. 2008 Mar;28(3):191-6.<br />

Liptay S, Schmid RM, Nabel EG, Nabel GJ.<br />

Transcriptional regulation <strong>of</strong> NF-κB2: evidence for κB-mediated positive and negative autoregulation. Mol Cell Biol.<br />

1994 Dec;14(12):7695-703.<br />

Liu GY, Essex A, Buchanan JT, Datta V, H<strong>of</strong>fman HM, Bastian JF, Fierer J, Nizet V.<br />

Staphylococcus aureus golden pigment impairs neutrophil killing and promotes virulence through its antioxidant<br />

activity. J Exp Med. 2005 Jul 18;202(2):209-15. Epub 2005 Jul 11.<br />

Liu J, Shue E, Ewalt KL, Schimmel P.<br />

A new gamma-interferon-inducible promoter and splice variants <strong>of</strong> an anti-angiogenic human tRNA synthetase.<br />

Nucleic Acids Res. 2004 Feb 2;32(2):719-27. Print 2004.<br />

Liu Z, Lu H, Jiang Z, Pastuszyn A, Hu CA.<br />

Apolipoprotein l6, a novel proapoptotic Bcl-2 homology 3-only protein, induces mitochondria-mediated apoptosis in<br />

cancer cells. Mol Cancer Res. 2005 Jan;3(1):21-31.<br />

Livermore DM.<br />

Antibiotic resistance in staphylococci. Int J Antimicrob Agents. 2000 Nov;16 Suppl 1:S3-10.<br />

Livermore DM.<br />

Has the era <strong>of</strong> untreatable infections arrived? J Antimicrob Chemother. 2009 Sep;64 Suppl 1:i29-36.<br />

Loetscher M, Gerber B, Loetscher P, Jones SA, Piali L, Clark-Lewis I, Baggiolini M, Moser B.<br />

Chemokine receptor specific for IP10 and Mig: structure, function, and expression in activated T-lymphocytes. J Exp<br />

Med. 1996 Sep 1;184(3):963-9.<br />

Loetscher P, Pellegrino A, Gong JH, Mattioli I, Loetscher M, Bardi G, Baggiolini M, Clark-Lewis I.<br />

The ligands <strong>of</strong> CXC chemokine receptor 3, I-TAC, Mig, and IP10, are natural antagonists for CCR3. J Biol Chem. 2001<br />

Feb 2;276(5):2986-91. Epub 2000 Nov 10.<br />

Lorenz U, Hüttinger C, Schäfer T, Ziebuhr W, Thiede A, Hacker J, Engelmann S, Hecker M, Ohlsen K.<br />

The alternative sigma factor sigma B <strong>of</strong> Staphylococcus aureus modulates virulence in experimental central venous<br />

catheter-related infections. Microbes Infect. 2008 Mar;10(3):217-23. Epub 2007 Nov 28.<br />

Loukissa A, Cardozo C, Altschuller-Felberg C, Nelson JE.<br />

Control <strong>of</strong> LMP7 expression in human endothelial cells <strong>by</strong> cytokines regulating cellular and humoral immunity.<br />

Cytokine. 2000 Sep;12(9):1326-30.<br />

Lowy FD.<br />

Staphylococcus aureus infections. N Engl J Med. 1998 Aug 20;339(8):520-32.<br />

Lowy FD.<br />

Is Staphylococcus aureus an intracellular <strong>pathogen</strong>? Trends Microbiol. 2000 Aug;8(8):341-3.<br />

Lundberg U.<br />

Stress hormones in health and illness: the roles <strong>of</strong> work and gender. Psychoneuroendocrinology. 2005 Nov;<br />

30(10):1017-21.<br />

221


Maren Depke<br />

References<br />

Luster AD, Unkeless JC, Ravetch JV.<br />

Gamma-interferon transcriptionally regulates an early-response gene containing homology to platelet proteins.<br />

Nature. 1985 Jun 20-26;315(6021):672-6.<br />

Luster AD.<br />

The role <strong>of</strong> chemokines in linking innate and adaptive immunity. Curr Opin Immunol. 2002 Feb;14(1):129-35.<br />

Ma J, Chen T, Mandelin J, Ceponis A, Miller NE, Hukkanen M, Ma GF, Konttinen YT.<br />

Regulation <strong>of</strong> macrophage activation. Cell Mol Life Sci. 2003 Nov;60(11):2334-46.<br />

Ma W, Lehner PJ, Cresswell P, Pober JS, Johnson DR.<br />

Interferon-gamma rapidly increases peptide transporter (TAP) subunit expression and peptide transport capacity in<br />

endothelial cells. J Biol Chem. 1997 Jun 27;272(26):16585-90.<br />

Mackey-Lawrence NM, Potter DE, Cerca N, Jefferson KK.<br />

Staphylococcus aureus immunodominant surface antigen B is a cell-surface associated nucleic acid binding protein.<br />

BMC Microbiol. 2009 Mar 26;9:61.<br />

MacMicking J, Xie QW, Nathan C.<br />

Nitric oxide and macrophage function. Annu Rev Immunol. 1997;15:323-50.<br />

MacMicking JD.<br />

IFN-inducible GTPases and immunity to intracellular <strong>pathogen</strong>s. Trends Immunol. 2004 Nov;25(11):601-9.<br />

Macpherson AJ, Martinic MM, Harris N.<br />

The functions <strong>of</strong> mucosal T cells in containing the indigenous commensal flora <strong>of</strong> the intestine. Cell Mol Life Sci.<br />

2002 Dec;59(12):2088-96.<br />

Maemura K, Zheng Q, Wada T, Ozaki M, Takao S, Aikou T, Bulkley GB, Klein AS, Sun Z.<br />

Reactive oxygen species are essential mediators in antigen presentation <strong>by</strong> Kupffer cells. Immunol Cell Biol. 2005<br />

Aug;83(4):336-43.<br />

Mak RH, Cheung W, Cone RD, Marks DL.<br />

Leptin and inflammation-associated cachexia in chronic kidney disease. Kidney Int. 2006 Mar;69(5):794-7.<br />

Mandell GL.<br />

Catalase, superoxide dismutase, and virulence <strong>of</strong> Staphylococcus aureus. In vitro and in vivo studies with emphasis<br />

on staphylococcal--leukocyte interaction. J Clin Invest. 1975 Mar;55(3):561-6.<br />

Mantovani A, Sica A, Sozzani S, Allavena P, Vecchi A, Locati M.<br />

The chemokine system in diverse forms <strong>of</strong> macrophage activation and polarization. Trends Immunol. 2004 Dec;<br />

25(12):677-86.<br />

Maresso AW, Schneewind O.<br />

Iron acquisition and transport in Staphylococcus aureus. Biometals. 2006 Apr;19(2):193-203.<br />

Marshall JH, Wilmoth GJ.<br />

Pigments <strong>of</strong> Staphylococcus aureus, a series <strong>of</strong> triterpenoid carotenoids. J Bacteriol. 1981 Sep;147(3):900-13.<br />

Martens S, Sabel K, Lange R, Uthaiah R, Wolf E, Howard JC.<br />

Mechanisms regulating the positioning <strong>of</strong> mouse p47 resistance GTPases LRG-47 and IIGP1 on cellular membranes:<br />

retargeting to plasma membrane induced <strong>by</strong> phagocytosis. J Immunol. 2004 Aug 15;173(4):2594-606.<br />

Martin G, Pilon A, Albert C, Vallé M, Hum DW, Fruchart JC, Najib J, Clavey V, Staels B.<br />

Comparison <strong>of</strong> expression and regulation <strong>of</strong> the high-density lipoprotein receptor SR-BI and the low-density<br />

lipoprotein receptor in human adrenocortical carcinoma NCI-H295 cells. Eur J Biochem. 1999 Apr;261(2):481-91.<br />

Matussek A, Strindhall J, Stark L, Rohde M, Geffers R, Buer J, Kihlström E, Lindgren PE, Löfgren S.<br />

Infection <strong>of</strong> human endothelial cells with Staphylococcus aureus induces transcription <strong>of</strong> genes encoding an innate<br />

immunity response. Scand J Immunol. 2005 Jun;61(6):536-44.<br />

McEwen BS.<br />

Protection and damage from acute and chronic stress: allostasis and allostatic overload and relevance to the<br />

pathophysiology <strong>of</strong> psychiatric disorders. Ann N Y Acad Sci. 2004 Dec;1032:1-7.<br />

McGuinness OP, Jacobs J, Moran C, Lacy B.<br />

Impact <strong>of</strong> infection on hepatic disposal <strong>of</strong> a peripheral glucose infusion in the conscious dog. Am J Physiol. 1995 Aug;<br />

269(2 Pt 1):E199-207.<br />

McGuinness OP, Snowden RT, Moran C, Neal DW, Fujiwara T, Cherrington AD.<br />

Impact <strong>of</strong> acute epinephrine removal on hepatic glucose metabolism during stress hormone infusion. Metabolism.<br />

1999 Jul;48(7):910-4.<br />

222


Maren Depke<br />

References<br />

McInnes IB, Gracie JA, Leung BP, Wei XQ, Liew FY.<br />

Interleukin 18: a pleiotropic participant in chronic inflammation. Immunol Today. 2000 Jul;21(7):312-5.<br />

McNamara PJ, Proctor RA.<br />

Staphylococcus aureus small colony variants, electron transport and persistent infections. Int J Antimicrob Agents.<br />

2000 Mar;14(2):117-22.<br />

Menestrina G, Serra MD, Prévost G.<br />

Mode <strong>of</strong> action <strong>of</strong> beta-barrel pore-forming toxins <strong>of</strong> the staphylococcal alpha-hemolysin family. Toxicon. 2001 Nov;<br />

39(11):1661-72.<br />

Miao L, St Clair DK.<br />

Regulation <strong>of</strong> superoxide dismutase genes: implications in disease. Free Radic Biol Med. 2009 Aug 15;47(4):344-56.<br />

Epub 2009 May 25.<br />

Mills CD, Kincaid K, Alt JM, Heilman MJ, Hill AM.<br />

M-1/M-2 macrophages and the Th1/Th2 paradigm. J Immunol. 2000 Jun 15;164(12):6166-73.<br />

Mittelstadt PR, Ashwell JD.<br />

Inhibition <strong>of</strong> AP-1 <strong>by</strong> the glucocorticoid-inducible protein GILZ. J Biol Chem. 2001 Aug 3;276(31):29603-10. Epub<br />

2001 Jun 7.<br />

Mizock BA.<br />

Alterations in carbohydrate metabolism during stress: a review <strong>of</strong> the literature. Am J Med. 1995 Jan;98(1):75-84.<br />

Mold C, Du Clos TW.<br />

C-reactive protein increases cytokine responses to Streptococcus pneumoniae through <strong>interactions</strong> with Fcγ<br />

receptors. J Immunol. 2006 Jun 15;176(12):7598-604.<br />

Moore KW, de Waal Malefyt R, C<strong>of</strong>fman RL, O'Garra A.<br />

Interleukin-10 and the interleukin-10 receptor. Annu Rev Immunol. 2001;19:683-765.<br />

Moreilhon C, Gras D, Hologne C, Bajolet O, Cottrez F, Magnone V, Merten M, Groux H, Puchelle E, Barbry P.<br />

Live Staphylococcus aureus and bacterial soluble factors induce different transcriptional responses in human airway<br />

cells. Physiol Genomics. 2005 Feb 10;20(3):244-55. Epub 2004 Dec 14.<br />

Morgan ET.<br />

Regulation <strong>of</strong> cytochrome p450 <strong>by</strong> inflammatory mediators: why and how? Drug Metab Dispos. 2001 Mar;<br />

29(3):207-12.<br />

Mori M, Gotoh T.<br />

Arginine metabolic enzymes, nitric oxide and infection. J Nutr. 2004 Oct;134(10 Suppl):2820S-2825.<br />

Morikawa K, Inose Y, Okamura H, Maruyama A, Hayashi H, Takeyasu K, Ohta T.<br />

A new staphylococcal sigma factor in the conserved gene cassette: functional significance and implication for the<br />

evolutionary processes. Genes Cells. 2003 Aug;8(8):699-712.<br />

Morley JE, Thomas DR, Wilson MM.<br />

Cachexia: pathophysiology and clinical relevance. Am J Clin Nutr. 2006 Apr;83(4):735-43.<br />

Mosser DM.<br />

The many faces <strong>of</strong> macrophage activation. J Leukoc Biol. 2003 Feb;73(2):209-12.<br />

Müller A, Heseler K, Schmidt SK, Spekker K, Mackenzie CR, Däubener W.<br />

The missing link between indoleamine 2,3-dioxygenase mediated antibacterial and immunoregulatory effects. J Cell<br />

Mol Med. 2009 Jun;13(6):1125-35. Epub 2009 Oct 13.<br />

Munn DH, Sharma MD, Baban B, Harding HP, Zhang Y, Ron D, Mellor AL.<br />

GCN2 kinase in T cells mediates proliferative arrest and anergy induction in response to indoleamine 2,3-<br />

dioxygenase. Immunity. 2005 May;22(5):633-42.<br />

Murdoch C, Finn A.<br />

Chemokine receptors and their role in inflammation and infectious diseases. Blood. 2000 May 15;95(10):3032-43.<br />

Murray HW, Szuro-Sudol A, Wellner D, Oca MJ, Granger AM, Lib<strong>by</strong> DM, Rothermel CD, Rubin BY.<br />

Role <strong>of</strong> tryptophan degradation in respiratory burst-independent antimicrobial activity <strong>of</strong> gamma interferonstimulated<br />

human macrophages. Infect Immun. 1989 Mar;57(3):845-9.<br />

Murray MF.<br />

Tryptophan depletion and HIV infection: a metabolic link to <strong>pathogen</strong>esis. Lancet Infect Dis. 2003 Oct;3(10):644-52.<br />

223


Maren Depke<br />

References<br />

Namiki S, Nakamura T, Oshima S, Yamazaki M, Sekine Y, Tsuchiya K, Okamoto R, Kanai T, Watanabe M.<br />

IRF-1 mediates upregulation <strong>of</strong> LMP7 <strong>by</strong> IFN-γ and concerted expression <strong>of</strong> immunosubunits <strong>of</strong> the proteasome.<br />

FEBS Lett. 2005 May 23;579(13):2781-7. Epub 2005 Apr 20.<br />

Narui K, Noguchi N, Saito A, Kakimi K, Motomura N, Kubo K, Takamoto S, Sasatsu M.<br />

Anti-infectious activity <strong>of</strong> tryptophan metabolites in the L-tryptophan-L-kynurenine pathway. Biol Pharm Bull. 2009<br />

Jan;32(1):41-4.<br />

Nemoto N, Sakurai J.<br />

Glucocorticoid and sex hormones as activating or modulating factors for expression <strong>of</strong> Cyp2b-9 and Cyp2b-10 in the<br />

mouse liver and hepatocytes. Arch Biochem Biophys. 1995 May 10;319(1):286-92.<br />

Newton R, Kuitert LM, Bergmann M, Adcock IM, Barnes PJ.<br />

Evidence for involvement <strong>of</strong> NF-κB in the transcriptional control <strong>of</strong> COX-2 gene expression <strong>by</strong> IL-1β. Biochem<br />

Biophys Res Commun. 1997 Aug 8;237(1):28-32.<br />

Nguyen VT, Kamio Y, Higuchi H.<br />

Single-molecule imaging <strong>of</strong> cooperative assembly <strong>of</strong> γ-hemolysin on erythrocyte membranes. EMBO J. 2003 Oct 1;<br />

22(19):4968-79.<br />

Nicholas RO, Li T, McDevitt D, Marra A, Sucoloski S, Demarsh PL, Gentry DR.<br />

Isolation and <strong>characterization</strong> <strong>of</strong> a sigB deletion mutant <strong>of</strong> Staphylococcus aureus. Infect Immun. 1999 Jul;<br />

67(7):3667-9.<br />

Nicolas P, Leduc A, Robin S, Rasmussen S, Jarmer H, Bessières P.<br />

Transcriptional landscape estimation from tiling array data using a model <strong>of</strong> signal shift and drift. Bioinformatics.<br />

2009 Sep 15;25(18):2341-7. Epub 2009 Jun 26.<br />

Nilsson IM, Lee JC, Bremell T, Rydén C, Tarkowski A.<br />

The role <strong>of</strong> staphylococcal polysaccharide microcapsule expression in septicemia and septic arthritis. Infect Immun.<br />

1997 Oct;65(10):4216-21.<br />

Novick D, Kim SH, Fantuzzi G, Reznikov LL, Dinarello CA, Rubinstein M.<br />

Interleukin-18 binding protein: a novel modulator <strong>of</strong> the Th1 cytokine response. Immunity. 1999 Jan;10(1):127-36.<br />

Novick RP.<br />

Mobile genetic elements and bacterial toxinoses: the superantigen-encoding <strong>pathogen</strong>icity islands <strong>of</strong><br />

Staphylococcus aureus. Plasmid. 2003 Mar;49(2):93-105.<br />

Ogimoto K, Harris MK Jr, Wisse BE.<br />

MyD88 is a key mediator <strong>of</strong> anorexia, but not weight loss, induced <strong>by</strong> lipopolysaccharide and interleukin-1β.<br />

Endocrinology. 2006 Sep;147(9):4445-53. Epub 2006 Jun 15.<br />

Ordway RW, Singer JJ, Walsh JV Jr.<br />

Direct regulation <strong>of</strong> ion channels <strong>by</strong> fatty acids. Trends Neurosci. 1991 Mar;14(3):96-100.<br />

Otani A, Slike BM, Dorrell MI, Hood J, Kinder K, Ewalt KL, Cheresh D, Schimmel P, Friedlander M.<br />

A fragment <strong>of</strong> human TrpRS as a potent antagonist <strong>of</strong> ocular angiogenesis. Proc Natl Acad Sci U S A. 2002 Jan 8;<br />

99(1):178-83. Epub 2002 Jan 2.<br />

Ott M, Gogvadze V, Orrenius S, Zhivotovsky B.<br />

Mitochondria, oxidative stress and cell death. Apoptosis. 2007 May;12(5):913-22.<br />

Otter JA, French GL.<br />

Molecular epidemiology <strong>of</strong> community-associated meticillin-resistant Staphylococcus aureus in Europe. Lancet<br />

Infect Dis. 2010 Apr;10(4):227-39.<br />

Page NM, Butlin DJ, Lomthaisong K, Lowry PJ.<br />

The human apolipoprotein L gene cluster: identification, classification, and sites <strong>of</strong> distribution. Genomics. 2001<br />

May 15;74(1):71-8.<br />

Pagels M, Fuchs S, Pané-Farré J, Kohler C, Menschner L, Hecker M, McNamarra PJ, Bauer MC, von Wachenfeldt C,<br />

Liebeke M, Lalk M, Sander G, von Eiff C, Proctor RA, Engelmann S.<br />

Redox sensing <strong>by</strong> a Rex-family repressor is involved in the regulation <strong>of</strong> anaerobic gene expression in Staphylococcus<br />

aureus. Mol Microbiol. 2010 Jun 1;76(5):1142-61. Epub 2010 Mar 30.<br />

Park JB.<br />

Phagocytosis induces superoxide formation and apoptosis in macrophages. Exp Mol Med. 2003 Oct 31;35(5):325-35.<br />

224


Maren Depke<br />

References<br />

Park JY, Pillinger MH, Abramson SB.<br />

Prostaglandin E 2 synthesis and secretion: the role <strong>of</strong> PGE 2 synthases. Clin Immunol. 2006 Jun;119(3):229-40. Epub<br />

2006 Mar 15.<br />

Pasini E, Aquilani R, Dioguardi FS.<br />

Amino acids: chemistry and metabolism in normal and hypercatabolic states. Am J Cardiol. 2004 Apr 22;93(8A):3A-<br />

5A.<br />

Pasparakis M, Alexopoulou L, Episkopou V, Kollias G.<br />

Immune and inflammatory responses in TNFα-deficient mice: a critical requirement for TNFα in the formation <strong>of</strong><br />

primary B cell follicles, follicular dendritic cell networks and germinal centers, and in the maturation <strong>of</strong> the humoral<br />

immune response. J Exp Med. 1996 Oct 1;184(4):1397-411.<br />

Patti JM, Allen BL, McGavin MJ, Höök M.<br />

MSCRAMM-mediated adherence <strong>of</strong> microorganisms to <strong>host</strong> tissues. Annu Rev Microbiol. 1994;48:585-617.<br />

Peacock SJ, Foster TJ, Cameron BJ, Berendt AR.<br />

Bacterial fibronectin-binding proteins and endothelial cell surface fibronectin mediate adherence <strong>of</strong> Staphylococcus<br />

aureus to resting human endothelial cells. Microbiology. 1999 Dec;145 (Pt 12):3477-86.<br />

Peacock SJ, Day NP, Thomas MG, Berendt AR, Foster TJ.<br />

Clinical isolates <strong>of</strong> Staphylococcus aureus exhibit diversity in fnb genes and adhesion to human fibronectin. J Infect.<br />

2000 Jul;41(1):23-31.<br />

Pereira CA, Modolell M, Frey JR, Lefkovits I.<br />

Gene expression in IFN-gamma-activated murine macrophages. Braz J Med Biol Res. 2004 Dec;37(12):1795-809.<br />

Epub 2004 Nov 17.<br />

Pérez-Morga D, Vanhollebeke B, Paturiaux-Hanocq F, Nolan DP, Lins L, Homblé F, Vanhamme L, Tebabi P, Pays A,<br />

Poelvoorde P, Jacquet A, Brasseur R, Pays E.<br />

Apolipoprotein L-I promotes trypanosome lysis <strong>by</strong> forming pores in lysosomal membranes. Science. 2005 Jul 15;<br />

309(5733):469-72.<br />

Périchon B, Courvalin P.<br />

VanA-type vancomycin-resistant Staphylococcus aureus. Antimicrob Agents Chemother. 2009 Nov;53(11):4580-7.<br />

Epub 2009 Jun 8.<br />

Peschel A, Otto M, Jack RW, Kalbacher H, Jung G, Götz F.<br />

Inactivation <strong>of</strong> the dlt operon in Staphylococcus aureus confers sensitivity to defensins, protegrins, and other<br />

antimicrobial peptides. J Biol Chem. 1999 Mar 26;274(13):8405-10.<br />

Peschel A, Jack RW, Otto M, Collins LV, Staubitz P, Nicholson G, Kalbacher H, Nieuwenhuizen WF, Jung G, Tarkowski A,<br />

van Kessel KP, van Strijp JA.<br />

Staphylococcus aureus resistance to human defensins and evasion <strong>of</strong> neutrophil killing via the novel virulence factor<br />

MprF is based on modification <strong>of</strong> membrane lipids with l-lysine. J Exp Med. 2001 May 7;193(9):1067-76.<br />

Peterson ML, Ault K, Kremer MJ, Klingelhutz AJ, Davis CC, Squier CA, Schlievert PM.<br />

The innate immune system is activated <strong>by</strong> stimulation <strong>of</strong> vaginal epithelial cells with Staphylococcus aureus and<br />

toxic shock syndrome toxin 1. Infect Immun. 2005 Apr;73(4):2164-74.<br />

Peterson PK, Chao CC, Molitor T, Murtaugh M, Strgar F, Sharp BM.<br />

Stress and <strong>pathogen</strong>esis <strong>of</strong> infectious disease. Rev Infect Dis. 1991 Jul-Aug;13(4):710-20.<br />

Petkovic V, Moghini C, Paoletti S, Uguccioni M, Gerber B.<br />

I-TAC/CXCL11 is a natural antagonist for CCR5. J Leukoc Biol. 2004 Sep;76(3):701-8. Epub 2004 Jun 3.<br />

Petri B, Bixel MG.<br />

Molecular events during leukocyte diapedesis. FEBS J. 2006 Oct;273(19):4399-407. Epub 2006 Sep 11.<br />

Pfefferkorn ER.<br />

Interferon γ blocks the growth <strong>of</strong> Toxoplasma gondii in human fibroblasts <strong>by</strong> inducing the <strong>host</strong> cells to degrade<br />

tryptophan. Proc Natl Acad Sci U S A. 1984 Feb;81(3):908-12.<br />

Plata K, Rosato AE, Wegrzyn G.<br />

Staphylococcus aureus as an infectious agent: overview <strong>of</strong> biochemistry and molecular genetics <strong>of</strong> its <strong>pathogen</strong>icity.<br />

Acta Biochim Pol. 2009;56(4):597-612. Epub 2009 Dec 11.<br />

Popov A, Schultze JL.<br />

IDO-expressing regulatory dendritic cells in cancer and chronic infection. J Mol Med. 2008 Feb;86(2):145-60. Epub<br />

2007 Sep 18.<br />

225


Maren Depke<br />

References<br />

Puccetti P.<br />

On watching the watchers: IDO and type I/II IFN. Eur J Immunol. 2007 Apr;37(4):876-9.<br />

Qu D, Wang Y, Esmon NL, Esmon CT.<br />

Regulated endothelial protein C receptor shedding is mediated <strong>by</strong> tumor necrosis factor-alpha converting<br />

enzyme/ADAM17. J Thromb Haemost. 2007 Feb;5(2):395-402. Epub 2006 Dec 7.<br />

Que YA, François P, Haefliger JA, Entenza JM, Vaudaux P, Moreillon P.<br />

Reassessing the role <strong>of</strong> Staphylococcus aureus clumping factor and fibronectin-binding protein <strong>by</strong> expression in<br />

Lactococcus lactis. Infect Immun. 2001 Oct;69(10):6296-302.<br />

Rachid S, Ohlsen K, Wallner U, Hacker J, Hecker M, Ziebuhr W.<br />

Alternative transcription factor σ B is involved in regulation <strong>of</strong> bi<strong>of</strong>ilm expression in a Staphylococcus aureus mucosal<br />

isolate. J Bacteriol. 2000 Dec;182(23):6824-6.<br />

Rahman MM, McFadden G.<br />

Modulation <strong>of</strong> tumor necrosis factor <strong>by</strong> microbial <strong>pathogen</strong>s. PLoS Pathog. 2006 Feb;2(2):e4.<br />

Rampone H, Martínez GL, Giraudo AT, Calzolari A, Nagel R.<br />

In vivo expression <strong>of</strong> exoprotein synthesis with a Sae mutant <strong>of</strong> Staphylococcus aureus. Can J Vet Res. 1996 Jul;<br />

60(3):237-40.<br />

Rasmussen S, Nielsen HB, Jarmer H.<br />

The transcriptionally active regions in the genome <strong>of</strong> Bacillus subtilis. Mol Microbiol. 2009 Sep;73(6):1043-57. Epub<br />

2009 Aug 4.<br />

Recsei PA, Gruss AD, Novick RP.<br />

Cloning, sequence, and expression <strong>of</strong> the lysostaphin gene from Staphylococcus simulans. Proc Natl Acad Sci U S A.<br />

1987 Mar;84(5):1127-31.<br />

Ricart-Jané D, Rodríguez-Sureda V, Benavides A, Peinado-Onsurbe J, López-Tejero MD, Llobera M.<br />

Immobilization stress alters intermediate metabolism and circulating lipoproteins in the rat. Metabolism. 2002 Jul;<br />

51(7):925-31.<br />

Ricquier D, Bouillaud F.<br />

Mitochondrial uncoupling proteins: from mitochondria to the regulation <strong>of</strong> energy balance. J Physiol. 2000 Nov 15;<br />

529 Pt 1:3-10.<br />

Riegert P, Wanner V, Bahram S.<br />

Genomics, is<strong>of</strong>orms, expression, and phylogeny <strong>of</strong> the MHC class I-related MR1 gene. J Immunol. 1998 Oct 15;<br />

161(8):4066-77.<br />

Riehle KJ, Campbell JS, McMahan RS, Johnson MM, Beyer RP, Bammler TK, Fausto N.<br />

Regulation <strong>of</strong> liver regeneration and hepatocarcinogenesis <strong>by</strong> suppressor <strong>of</strong> cytokine signaling 3. J Exp Med. 2008<br />

Jan 21;205(1):91-103. Epub 2007 Dec 24.<br />

Rivett AJ, Bose S, Brooks P, Broadfoot KI.<br />

Regulation <strong>of</strong> proteasome complexes <strong>by</strong> gamma-interferon and phosphorylation. Biochimie. 2001 Mar-Apr;83(3-4):<br />

363-6.<br />

Rizki G, Arnaboldi L, Gabrielli B, Yan J, Lee GS, Ng RK, Turner SM, Badger TM, Pitas RE, Maher JJ.<br />

Mice fed a lipogenic methionine-choline-deficient diet develop hypermetabolism coincident with hepatic<br />

suppression <strong>of</strong> SCD-1. J Lipid Res. 2006 Oct;47(10):2280-90. Epub 2006 Jul 8.<br />

Robinson JM.<br />

Phagocytic leukocytes and reactive oxygen species. Histochem Cell Biol. 2009 Apr;131(4):465-9. Epub 2009 Feb 18.<br />

Rogasch K, Rühmling V, Pané-Farré J, Höper D, Weinberg C, Fuchs S, Schmudde M, Bröker BM, Wolz C, Hecker M,<br />

Engelmann S.<br />

Influence <strong>of</strong> the two-component system SaeRS on global gene expression in two different Staphylococcus aureus<br />

strains. J Bacteriol. 2006 Nov;188(22):7742-58.<br />

Rooijakkers SH, van Wamel WJ, Ruyken M, van Kessel KP, van Strijp JA.<br />

Anti-opsonic properties <strong>of</strong> staphylokinase. Microbes Infect. 2005 Mar;7(3):476-84. Epub 2005 Feb 26.<br />

Rooijakkers SH, Ruyken M, Roos A, Daha MR, Presanis JS, Sim RB, van Wamel WJ, van Kessel KP, van Strijp JA.<br />

Immune evasion <strong>by</strong> a staphylococcal complement inhibitor that acts on C3 convertases. Nat Immunol. 2005 Sep;<br />

6(9):920-7. Epub 2005 Aug 7.<br />

226


Maren Depke<br />

References<br />

Roudkenar MH, Kuwahara Y, Baba T, Roushandeh AM, Ebishima S, Abe S, Ohkubo Y, Fukumoto M.<br />

Oxidative stress induced lipocalin 2 gene expression: addressing its expression under the harmful conditions.<br />

J Radiat Res (Tokyo). 2007 Jan;48(1):39-44. Epub 2007 Jan 16.<br />

Rubin BY, Anderson SL, Xing L, Powell RJ, Tate WP.<br />

Interferon induces tryptophanyl-tRNA synthetase expression in human fibroblasts. J Biol Chem. 1991 Dec 25;<br />

266(36):24245-8.<br />

Saito A, Motomura N, Kakimi K, Narui K, Noguchi N, Sasatsu M, Kubo K, Koezuka Y, Takai D, Ueha S, Takamoto S.<br />

Vascular allografts are resistant to methicillin-resistant Staphylococcus aureus through indoleamine 2,3-dioxygenase<br />

in a murine model. J Thorac Cardiovasc Surg. 2008 Jul;136(1):159-67. Epub 2008 May 22.<br />

Sarvas M, Harwood CR, Bron S, van Dijl JM.<br />

Post-translocational folding <strong>of</strong> secretory proteins in Gram-positive bacteria. Biochim Biophys Acta. 2004 Nov 11;<br />

1694(1-3):311-27.<br />

Schiffer R, Baron J, Dagtekin G, Jahnen-Dechent W, Zwadlo-Klarwasser G.<br />

Differential regulation <strong>of</strong> the expression <strong>of</strong> transporters associated with antigen processing, TAP1 and TAP2, <strong>by</strong><br />

cytokines and lipopolysaccharide in primary human macrophages. Inflamm Res. 2002 Aug;51(8):403-8.<br />

Schindler CA, Schuhardt VT.<br />

Lysostaphin: A new bacteriolytic agent for the Staphylococcus. Proc Natl Acad Sci U S A. 1964 Mar;51:414-21.<br />

Schmidt F, Scharf SS, Hildebrandt P, Burian M, Bernhardt J, Dhople V, Kalinka J, Gutjahr M, Hammer E, Völker U.<br />

Time resolved quantitative proteome pr<strong>of</strong>iling <strong>of</strong> <strong>host</strong>-<strong>pathogen</strong> <strong>interactions</strong>: The response <strong>of</strong> Staphylococcus<br />

aureus RN1HG to internalisation <strong>by</strong> human airway epithelial cells. Proteomics. 2010 Aug;10(15):2801-11.<br />

Schröder JM.<br />

Epithelial antimicrobial peptides: innate local <strong>host</strong> response elements. Cell Mol Life Sci. 1999 Oct 1;56(1-2):32-46.<br />

Schroder K, Hertzog PJ, Ravasi T, Hume DA.<br />

Interferon-γ: an overview <strong>of</strong> signals, mechanisms and functions. J Leukoc Biol. 2004 Feb;75(2):163-89. Epub 2003<br />

Oct 2.<br />

Schwarz-Linek U, Höök M, Potts JR.<br />

The molecular basis <strong>of</strong> fibronectin-mediated bacterial adherence to <strong>host</strong> cells. Mol Microbiol. 2004 May;52(3):<br />

631-41.<br />

Scriba TJ, Sierro S, Brown EL, Phillips RE, Sewell AK, Massey RC.<br />

The Staphyloccous aureus Eap protein activates expression <strong>of</strong> proinflammatory cytokines. Infect Immun. 2008 May;<br />

76(5):2164-8. Epub 2008 Mar 10.<br />

Sekkaï D, Guittet O, Lemaire G, Tenu JP, Lepoivre M.<br />

Inhibition <strong>of</strong> nitric oxide synthase expression and activity in macrophages <strong>by</strong> 3-hydroxyanthranilic acid, a tryptophan<br />

metabolite. Arch Biochem Biophys. 1997 Apr 1;340(1):117-23.<br />

Senn MM, Giachino P, Homerova D, Steinhuber A, Strassner J, Kormanec J, Flückiger U, Berger-Bächi B, Bisch<strong>of</strong>f M.<br />

Molecular analysis and organization <strong>of</strong> the sigma B operon in Staphylococcus aureus. J Bacteriol. 2005 Dec;<br />

187(23):8006-19.<br />

Sharpe AH, Wherry EJ, Ahmed R, Freeman GJ.<br />

The function <strong>of</strong> programmed cell death 1 and its ligands in regulating autoimmunity and infection. Nat Immunol.<br />

2007 Mar;8(3):239-45.<br />

Shaw LN, Aish J, Davenport JE, Brown MC, Lithgow JK, Simmonite K, Crossley H, Travis J, Potempa J, Foster SJ.<br />

Investigations into σ B -modulated regulatory pathways governing extracellular virulence determinant production in<br />

Staphylococcus aureus. J Bacteriol. 2006 Sep;188(17):6070-80.<br />

Shaw LN, Lindholm C, Prajsnar TK, Miller HK, Brown MC, Golonka E, Stewart GC, Tarkowski A, Potempa J.<br />

Identification and <strong>characterization</strong> <strong>of</strong> σ S , a novel component <strong>of</strong> the Staphylococcus aureus stress and virulence<br />

responses. PLoS One. 2008;3(12):e3844. Epub 2008 Dec 3.<br />

Shenoy AR, Kim BH, Choi HP, Matsuzawa T, Tiwari S, MacMicking JD.<br />

Emerging themes in IFN-γ-induced macrophage immunity <strong>by</strong> the p47 and p65 GTPase families. Immunobiology.<br />

2007;212(9-10):771-84. Epub 2007 Nov 28.<br />

Sieprawska-Lupa M, Mydel P, Krawczyk K, Wójcik K, Puklo M, Lupa B, Suder P, Silberring J, Reed M, Pohl J, Shafer W,<br />

McAleese F, Foster T, Travis J, Potempa J.<br />

Degradation <strong>of</strong> human antimicrobial peptide LL-37 <strong>by</strong> Staphylococcus aureus-derived proteinases. Antimicrob<br />

Agents Chemother. 2004 Dec;48(12):4673-9.<br />

227


Maren Depke<br />

References<br />

Siewert E, Bort R, Kluge R, Heinrich PC, Castell J, Jover R.<br />

Hepatic cytochrome P450 down-regulation during aseptic inflammation in the mouse is interleukin 6 dependent.<br />

Hepatology. 2000 Jul;32(1):49-55.<br />

Silence K, Hartmann M, Gührs KH, Gase A, Schlott B, Collen D, Lijnen HR.<br />

Structure-function relationships in staphylokinase as revealed <strong>by</strong> "clustered charge to alanine" mutagenesis. J Biol<br />

Chem. 1995 Nov 10;270(45):27192-8.<br />

Sinars CR, Cheung-Flynn J, Rimerman RA, Scammell JG, Smith DF, Clardy J.<br />

Structure <strong>of</strong> the large FK506-binding protein FKBP51, an Hsp90-binding protein and a component <strong>of</strong> steroid<br />

receptor complexes. Proc Natl Acad Sci U S A. 2003 Feb 4;100(3):868-73. Epub 2003 Jan 21.<br />

Singh VK, Moskovitz J.<br />

Multiple methionine sulfoxide reductase genes in Staphylococcus aureus: expression <strong>of</strong> activity and roles in<br />

tolerance <strong>of</strong> oxidative stress. Microbiology. 2003 Oct;149(Pt 10):2739-47.<br />

Sinha B, François PP, Nüsse O, Foti M, Hartford OM, Vaudaux P, Foster TJ, Lew DP, Herrmann M, Krause KH.<br />

Fibronectin-binding protein acts as Staphylococcus aureus invasin via fibronectin bridging to integrin alpha5beta1.<br />

Cell Microbiol. 1999 Sep;1(2):101-17.<br />

Sinha B, Herrmann M, Krause KH.<br />

Is Staphylococcus aureus an intracellular <strong>pathogen</strong>? Trends Microbiol. 2000 Aug;8(8):343-4. Response to Lowy, 2000.<br />

Siqueira JA, Speeg-Schatz C, Freitas FI, Sahel J, Monteil H, Prévost G.<br />

Channel-forming leucotoxins from Staphylococcus aureus cause severe inflammatory reactions in a rabbit eye<br />

model. J Med Microbiol. 1997 Jun;46(6):486-94.<br />

Smith EE, Malik HS.<br />

The apolipoprotein L family <strong>of</strong> programmed cell death and immunity genes rapidly evolved in primates at discrete<br />

sites <strong>of</strong> <strong>host</strong>-<strong>pathogen</strong> <strong>interactions</strong>. Genome Res. 2009 May;19(5):850-8. Epub 2009 Mar 19.<br />

Sobke AC, Selimovic D, Orlova V, Hassan M, Chavakis T, Athanasopoulos AN, Schubert U, Hussain M, Thiel G,<br />

Preissner KT, Herrmann M.<br />

The extracellular adherence protein from Staphylococcus aureus abrogates angiogenic responses <strong>of</strong> endothelial cells<br />

<strong>by</strong> blocking Ras activation. FASEB J. 2006 Dec;20(14):2621-3. Epub 2006 Oct 31.<br />

Souba WW, Smith RJ, Wilmore DW.<br />

Effects <strong>of</strong> glucocorticoids on glutamine metabolism in visceral organs. Metabolism. 1985 May;34(5):450-6.<br />

Speciale C, Hares K, Schwarcz R, Brookes N.<br />

High-affinity uptake <strong>of</strong> L-kynurenine <strong>by</strong> a Na + -independent transporter <strong>of</strong> neutral amino acids in astrocytes.<br />

J Neurosci. 1989 Jun;9(6):2066-72.<br />

Spencer AG, Woods JW, Arakawa T, Singer II, Smith WL.<br />

Subcellular localization <strong>of</strong> prostaglandin endoperoxide H synthases-1 and -2 <strong>by</strong> immunoelectron microscopy. J Biol<br />

Chem. 1998 Apr 17;273(16):9886-93.<br />

Staeheli P, Prochazka M, Steigmeier PA, Haller O.<br />

Genetic control <strong>of</strong> interferon action: mouse strain distribution and inheritance <strong>of</strong> an induced protein with<br />

guanylate-binding property. Virology. 1984 Aug;137(1):135-42.<br />

Stapleton PD, Taylor PW.<br />

Methicillin resistance in Staphylococcus aureus: mechanisms and modulation. Sci Prog. 2002;85(Pt 1):57-72.<br />

Steer SA, Corbett JA.<br />

The role and regulation <strong>of</strong> COX-2 during viral infection. Viral Immunol. 2003;16(4):447-60.<br />

Steinhuber A, Goerke C, Bayer MG, Döring G, Wolz C.<br />

Molecular architecture <strong>of</strong> the regulatory Locus sae <strong>of</strong> Staphylococcus aureus and its impact on expression <strong>of</strong><br />

virulence factors. J Bacteriol. 2003 Nov;185(21):6278-86.<br />

Strehl B, Seifert U, Krüger E, Heink S, Kuckelkorn U, Kloetzel PM.<br />

Interferon-gamma, the functional plasticity <strong>of</strong> the ubiquitin-proteasome system, and MHC class I antigen<br />

processing. Immunol Rev. 2005 Oct;207:19-30.<br />

Sugita H, Kaneki M, Tokunaga E, Sugita M, Koike C, Yasuhara S, Tompkins RG, Martyn JA.<br />

Inducible nitric oxide synthase plays a role in LPS-induced hyperglycemia and insulin resistance. Am J Physiol<br />

Endocrinol Metab. 2002 Feb;282(2):E386-94.<br />

228


Maren Depke<br />

References<br />

Sun SC, Ganchi PA, Ballard DW, Greene WC.<br />

NF-κB controls expression <strong>of</strong> inhibitor IκBα: evidence for an inducible autoregulatory pathway. Science. 1993 Mar<br />

26;259(5103):1912-5.<br />

Swain MG.<br />

I. Stress and hepatic inflammation. Am J Physiol Gastrointest Liver Physiol. 2000 Dec;279(6):G1135-8.<br />

Swearingen KE, Loomis WP, Zheng M, Cookson BT, Dovichi NJ.<br />

Proteomic pr<strong>of</strong>iling <strong>of</strong> lipopolysaccharide-activated macrophages <strong>by</strong> isotope coded affinity tagging. J Proteome Res.<br />

2010 May 7;9(5):2412-21.<br />

Taub DD, Lloyd AR, Conlon K, Wang JM, Ortaldo JR, Harada A, Matsushima K, Kelvin DJ, Oppenheim JJ.<br />

Recombinant human interferon-inducible protein 10 is a chemoattractant for human monocytes and T lymphocytes<br />

and promotes T cell adhesion to endothelial cells. J Exp Med. 1993 Jun 1;177(6):1809-14.<br />

Taylor GA, Feng CG, Sher A.<br />

Control <strong>of</strong> IFN-γ-mediated <strong>host</strong> resistance to intracellular <strong>pathogen</strong>s <strong>by</strong> immunity-related GTPases (p47 GTPases).<br />

Microbes Infect. 2007 Nov-Dec;9(14-15):1644-51. Epub 2007 Sep 14.<br />

Taylor MW, Feng GS.<br />

Relationship between interferon-γ, indoleamine 2,3-dioxygenase, and tryptophan catabolism. FASEB J. 1991 Aug;<br />

5(11):2516-22.<br />

Thakker M, Park JS, Carey V, Lee JC.<br />

Staphylococcus aureus serotype 5 capsular polysaccharide is antiphagocytic and enhances bacterial virulence in a<br />

murine bacteremia model. Infect Immun. 1998 Nov;66(11):5183-9.<br />

Thomas SR, Mohr D, Stocker R.<br />

Nitric oxide inhibits indoleamine 2,3-dioxygenase activity in interferon-γ primed mononuclear phagocytes. J Biol<br />

Chem. 1994 May 20;269(20):14457-64.<br />

Tolstrup AB, Bejder A, Fleckner J, Justesen J.<br />

Transcriptional regulation <strong>of</strong> the interferon-gamma-inducible tryptophanyl-tRNA synthetase includes alternative<br />

splicing. J Biol Chem. 1995 Jan 6;270(1):397-403.<br />

Trinh DV, Zhu N, Farhang G, Kim BJ, Huxford T.<br />

The nuclear IκB protein IκBζ specifically binds NF-κB p50 homodimers and forms a ternary complex on κB DNA. J Mol<br />

Biol. 2008 May 23;379(1):122-35. Epub 2008 Apr 3.<br />

Tsatsanis C, Androulidaki A, Venihaki M, Margioris AN.<br />

Signalling networks regulating cyclooxygenase-2. Int J Biochem Cell Biol. 2006;38(10):1654-61. Epub 2006 Apr 25.<br />

Tseng CF, Lin CC, Huang HY, Liu HC, Mao SJ.<br />

Antioxidant role <strong>of</strong> human haptoglobin. Proteomics. 2004 Aug;4(8):2221-8.<br />

Tu L, Moriya C, Imai T, Ishida H, Tetsutani K, Duan X, Murata S, Tanaka K, Shimokawa C, Hisaeda H, Himeno K.<br />

Critical role for the immunoproteasome subunit LMP7 in the resistance <strong>of</strong> mice to Toxoplasma gondii infection.<br />

Eur J Immunol. 2009 Dec;39(12):3385-94.<br />

Uhlén M, Guss B, Nilsson B, Gatenbeck S, Philipson L, Lindberg M.<br />

Complete sequence <strong>of</strong> the staphylococcal gene encoding protein A. A gene evolved through multiple duplications.<br />

J Biol Chem. 1984 Feb 10;259(3):1695-702.<br />

Urieli-Shoval S, Meek RL, Hanson RH, Eriksen N, Benditt EP.<br />

Human serum amyloid A genes are expressed in monocyte/macrophage cell lines. Am J Pathol. 1994 Sep;<br />

145(3):650-60.<br />

Valderas MW, Hart ME.<br />

Identification and <strong>characterization</strong> <strong>of</strong> a second superoxide dismutase gene (sodM) from Staphylococcus aureus.<br />

J Bacteriol. 2001 Jun;183(11):3399-407.<br />

Valderas MW, Gatson JW, Wreyford N, Hart ME.<br />

The superoxide dismutase gene sodM is unique to Staphylococcus aureus: absence <strong>of</strong> sodM in coagulase-negative<br />

staphylococci. J Bacteriol. 2002 May;184(9):2465-72.<br />

Valle J, Toledo-Arana A, Berasain C, Ghigo JM, Amorena B, Penadés JR, Lasa I.<br />

SarA and not σ B is essential for bi<strong>of</strong>ilm development <strong>by</strong> Staphylococcus aureus. Mol Microbiol. 2003 May;<br />

48(4):1075-87.<br />

van Belkum A, Verkaik NJ, de Vogel CP, Boelens HA, Verveer J, Nouwen JL, Verbrugh HA, Wertheim HF.<br />

Reclassification <strong>of</strong> Staphylococcus aureus nasal carriage types. J Infect Dis. 2009 Jun 15;199(12):1820-6.<br />

229


Maren Depke<br />

References<br />

van den Akker EL, Nouwen JL, Melles DC, van Rossum EF, Koper JW, Uitterlinden AG, H<strong>of</strong>man A, Verbrugh HA,<br />

Pols HA, Lamberts SW, van Belkum A.<br />

Staphylococcus aureus nasal carriage is associated with glucocorticoid receptor gene polymorphisms. J Infect Dis.<br />

2006 Sep 15;194(6):814-8. Epub 2006 Aug 8.<br />

Van den Eynde BJ, Morel S.<br />

Differential processing <strong>of</strong> class-I-restricted epitopes <strong>by</strong> the standard proteasome and the immunoproteasome. Curr<br />

Opin Immunol. 2001 Apr;13(2):147-53.<br />

van Erp K, Dach K, Koch I, Heesemann J, H<strong>of</strong>fmann R.<br />

Role <strong>of</strong> strain differences on <strong>host</strong> resistance and the transcriptional response <strong>of</strong> macrophages to infection with<br />

Yersinia enterocolitica. Physiol Genomics. 2006 Mar 13;25(1):75-84. Epub 2005 Dec 13.<br />

van Oosten M, van Amersfoort ES, van Berkel TJ, Kuiper J.<br />

Scavenger receptor-like receptors for the binding <strong>of</strong> lipopolysaccharide and lipoteichoic acid to liver endothelial and<br />

Kupffer cells. J Endotoxin Res. 2001;7(5):381-4.<br />

van Waardenburg DA, Jansen TC, Vos GD, Buurman WA.<br />

Hyperglycemia in children with meningococcal sepsis and septic shock: the relation between plasma levels <strong>of</strong> insulin<br />

and inflammatory mediators. J Clin Endocrinol Metab. 2006 Oct;91(10):3916-21. Epub 2006 May 30.<br />

Vanhamme L, Paturiaux-Hanocq F, Poelvoorde P, Nolan DP, Lins L, Van Den Abbeele J, Pays A, Tebabi P, Van Xong H,<br />

Jacquet A, Moguilevsky N, Dieu M, Kane JP, De Baetselier P, Brasseur R, Pays E.<br />

Apolipoprotein L-I is the trypanosome lytic factor <strong>of</strong> human serum. Nature. 2003 Mar 6;422(6927):83-7.<br />

Vanhollebeke B, Pays E.<br />

The function <strong>of</strong> apolipoproteins L. Cell Mol Life Sci. 2006 Sep;63(17):1937-44.<br />

Vanhorebeek I, Van den Berghe G.<br />

Hormonal and metabolic strategies to attenuate catabolism in critically ill patients. Curr Opin Pharmacol. 2004<br />

Dec;4(6):621-8.<br />

Vilcek J.<br />

First demonstration <strong>of</strong> the role <strong>of</strong> TNF in the <strong>pathogen</strong>esis <strong>of</strong> disease. J Immunol. 2008 Jul 1;181(1):5-6.<br />

Viswanathan K, Dhabhar FS.<br />

Stress-induced enhancement <strong>of</strong> leukocyte trafficking into sites <strong>of</strong> surgery or immune activation. Proc Natl Acad Sci<br />

U S A. 2005 Apr 19;102(16):5808-13. Epub 2005 Apr 7.<br />

von Eiff C, Becker K, Machka K, Stammer H, Peters G.<br />

Nasal carriage as a source <strong>of</strong> Staphylococcus aureus bacteremia. Study Group. N Engl J Med. 2001 Jan 4;344(1):11-6.<br />

von Eiff C, Friedrich AW, Peters G, Becker K.<br />

Prevalence <strong>of</strong> genes encoding for members <strong>of</strong> the staphylococcal leukotoxin family among clinical isolates <strong>of</strong><br />

Staphylococcus aureus. Diagn Microbiol Infect Dis. 2004 Jul;49(3):157-62.<br />

von Eiff C, Peters G, Becker K.<br />

The small colony variant (SCV) concept – the role <strong>of</strong> staphylococcal SCVs in persistent infections. Injury. 2006<br />

May;37 Suppl 2:S26-33.<br />

Voyich JM, Braughton KR, Sturdevant DE, Whitney AR, Saïd-Salim B, Porcella SF, Long RD, Dorward DW, Gardner DJ,<br />

Kreiswirth BN, Musser JM, DeLeo FR.<br />

Insights into mechanisms used <strong>by</strong> Staphylococcus aureus to avoid destruction <strong>by</strong> human neutrophils. J Immunol.<br />

2005 Sep 15;175(6):3907-19.<br />

Wajant H, Pfizenmaier K, Scheurich P.<br />

Tumor necrosis factor signaling. Cell Death Differ. 2003 Jan;10(1):45-65.<br />

Wakasugi K, Slike BM, Hood J, Otani A, Ewalt KL, Friedlander M, Cheresh DA, Schimmel P.<br />

A human aminoacyl-tRNA synthetase as a regulator <strong>of</strong> angiogenesis. Proc Natl Acad Sci U S A. 2002 Jan 8;99(1):173-<br />

7. Epub 2002 Jan 2.<br />

Wan G, Zhaorigetu S, Liu Z, Kaini R, Jiang Z, Hu CA.<br />

Apolipoprotein L1, a novel Bcl-2 homology domain 3-only lipid-binding protein, induces autophagic cell death. J Biol<br />

Chem. 2008 Aug 1;283(31):21540-9. Epub 2008 May 26.<br />

Wang J, Maldonado MA.<br />

The ubiquitin-proteasome system and its role in inflammatory and autoimmune diseases. Cell Mol Immunol. 2006<br />

Aug;3(4):255-61.<br />

230


Maren Depke<br />

References<br />

Weekers F, Van Herck E, Coopmans W, Michalaki M, Bowers CY, Veldhuis JD, Van den Berghe G.<br />

A novel in vivo rabbit model <strong>of</strong> hypercatabolic critical illness reveals a biphasic neuroendocrine stress response.<br />

Endocrinology. 2002 Mar;143(3):764-74.<br />

Wertheim HF, Vos MC, Ott A, van Belkum A, Voss A, Kluytmans JA, van Keulen PH, Vandenbroucke-Grauls CM,<br />

Meester MH, Verbrugh HA.<br />

Risk and outcome <strong>of</strong> nosocomial Staphylococcus aureus bacteraemia in nasal carriers versus non-carriers. Lancet.<br />

2004 Aug 21-27;364(9435):703-5.<br />

Wertheim HF, Melles DC, Vos MC, van Leeuwen W, van Belkum A, Verbrugh HA, Nouwen JL.<br />

The role <strong>of</strong> nasal carriage in Staphylococcus aureus infections. Lancet Infect Dis. 2005 Dec;5(12):751-62.<br />

West NP, Pyne DB, Renshaw G, Cripps AW.<br />

Antimicrobial peptides and proteins, exercise and innate mucosal immunity. FEMS Immunol Med Microbiol. 2006<br />

Dec;48(3):293-304.<br />

White LC, Wright KL, Felix NJ, Ruffner H, Reis LF, Pine R, Ting JP.<br />

Regulation <strong>of</strong> LMP2 and TAP1 genes <strong>by</strong> IRF-1 explains the paucity <strong>of</strong> CD8 + T cells in IRF-1 -/- mice. Immunity. 1996<br />

Oct;5(4):365-76.<br />

Wiegard C, Frenzel C, Herkel J, Kallen KJ, Schmitt E, Lohse AW.<br />

Murine liver antigen presenting cells control suppressor activity <strong>of</strong> CD4+CD25+ regulatory T cells. Hepatology. 2005<br />

Jul;42(1):193-9.<br />

Wiekowski MT, Chen SC, Zalamea P, Wilburn BP, Kinsley DJ, Sharif WW, Jensen KK, Hedrick JA, Manfra D, Lira SA.<br />

Disruption <strong>of</strong> neutrophil migration in a conditional transgenic model: evidence for CXCR2 desensitization in vivo.<br />

J Immunol. 2001 Dec 15;167(12):7102-10.<br />

Wilmore DW.<br />

Metabolic response to severe surgical illness: overview. World J Surg. 2000 Jun;24(6):705-11.<br />

Wolz C, Geiger T, Goerke C.<br />

The synthesis and function <strong>of</strong> the alarmone (p)ppGpp in firmicutes. Int J Med Microbiol. 2010 Feb;300(2-3):142-7.<br />

Epub 2009 Sep 24.<br />

Wray CJ, Mammen JM, Hasselgren PO.<br />

Catabolic response to stress and potential benefits <strong>of</strong> nutrition support. Nutrition. 2002 Nov-Dec;18(11-12):971-7.<br />

Wrona D.<br />

Neural-immune <strong>interactions</strong>: an integrative view <strong>of</strong> the bidirectional relationship between the brain and immune<br />

systems. J Neuroimmunol. 2006 Mar;172(1-2):38-58. Epub 2006 Jan 10.<br />

Wu S, de Lencastre H, Tomasz A.<br />

Sigma-B, a putative operon encoding alternate sigma factor <strong>of</strong> Staphylococcus aureus RNA polymerase: molecular<br />

cloning and DNA sequencing. J Bacteriol. 1996 Oct;178(20):6036-42.<br />

Wu SW, de Lencastre H, Tomasz A.<br />

Recruitment <strong>of</strong> the mecA gene homologue <strong>of</strong> Staphylococcus sciuri into a resistance determinant and expression <strong>of</strong><br />

the resistant phenotype in Staphylococcus aureus. J Bacteriol. 2001 Apr;183(8):2417-24.<br />

Wurfel MM, Kunitake ST, Lichenstein H, Kane JP, Wright SD.<br />

Lipopolysaccharide (LPS)-binding protein is carried on lipoproteins and acts as a c<strong>of</strong>actor in the neutralization <strong>of</strong> LPS.<br />

J Exp Med. 1994 Sep 1;180(3):1025-35.<br />

Xie G, Bonner CA, Jensen RA.<br />

Dynamic diversity <strong>of</strong> the tryptophan pathway in chlamydiae: reductive evolution and a novel operon for tryptophan<br />

recapture. Genome Biol. 2002 Aug 29;3(9):research0051. Epub 2002 Aug 29.<br />

Xong HV, Vanhamme L, Chamekh M, Chimfwembe CE, Van Den Abbeele J, Pays A, Van Meirvenne N, Hamers R, De<br />

Baetselier P, Pays E.<br />

A VSG expression site-associated gene confers resistance to human serum in Trypanosoma rhodesiense. Cell. 1998<br />

Dec 11;95(6):839-46.<br />

Xue H, Wong JT.<br />

Interferon induction <strong>of</strong> human tryptophanyl-tRNA synthetase safeguards the synthesis <strong>of</strong> tryptophan-rich immunesystem<br />

proteins: a hypothesis. Gene. 1995 Nov 20;165(2):335-9.<br />

Yang D, Chen Q, Hoover DM, Staley P, Tucker KD, Lubkowski J, Oppenheim JJ.<br />

Many chemokines including CCL20/MIP-3alpha display antimicrobial activity. J Leukoc Biol. 2003 Sep;74(3):448-55.<br />

231


Maren Depke<br />

References<br />

Yang XL, Schimmel P, Ewalt KL.<br />

Relationship <strong>of</strong> two human tRNA synthetases used in cell signaling. Trends Biochem Sci. 2004 May;29(5):250-6.<br />

Yao L, Bengualid V, Berman JW, Lowy FD.<br />

Prevention <strong>of</strong> endothelial cell cytokine induction <strong>by</strong> a Staphylococcus aureus lipoprotein. FEMS Immunol Med<br />

Microbiol. 2000 Aug;28(4):301-5.<br />

Yoo JY, Desiderio S.<br />

Innate and acquired immunity intersect in a global view <strong>of</strong> the acute-phase response. Proc Natl Acad Sci U S A. 2003<br />

Feb 4;100(3):1157-62. Epub 2003 Jan 22.<br />

Zangar RC, Davydov DR, Verma S.<br />

Mechanisms that regulate production <strong>of</strong> reactive oxygen species <strong>by</strong> cytochrome P450. Toxicol Appl Pharmacol. 2004<br />

Sep 15;199(3):316-31.<br />

Zeitlin PL, Lu L, Rhim J, Cutting G, Stetten G, Kieffer KA, Craig R, Guggino WB.<br />

A cystic fibrosis bronchial epithelial cell line: immortalization <strong>by</strong> adeno-12-SV40 infection. Am J Respir Cell Mol Biol.<br />

1991 Apr;4(4):313-9.<br />

Zelante T, Fallarino F, Bistoni F, Puccetti P, Romani L.<br />

Indoleamine 2,3-dioxygenase in infection: the paradox <strong>of</strong> an evasive strategy that benefits the <strong>host</strong>. Microbes Infect.<br />

2009 Jan;11(1):133-41. Epub 2008 Oct 25.<br />

Zhang K, McClure JA, Elsayed S, Conly JM.<br />

Novel staphylococcal cassette chromosome mec type, tentatively designated type VIII, harboring class A mec and<br />

type 4 ccr gene complexes in a Canadian epidemic strain <strong>of</strong> methicillin-resistant Staphylococcus aureus. Antimicrob<br />

Agents Chemother. 2009 Feb;53(2):531-40. Epub 2008 Dec 8.<br />

Zhaorigetu S, Wan G, Kaini R, Jiang Z, Hu CA.<br />

ApoL1, a BH3-only lipid-binding protein, induces autophagic cell death. Autophagy. 2008 Nov 16;4(8):1079-82. Epub<br />

2008 Nov 23.<br />

Ziebandt AK, Becher D, Ohlsen K, Hacker J, Hecker M, Engelmann S.<br />

The influence <strong>of</strong> agr and σ B in growth phase dependent regulation <strong>of</strong> virulence factors in Staphylococcus aureus.<br />

Proteomics. 2004 Oct;4(10):3034-47.<br />

Zocco MA, Carloni E, Pescatori M, Saulnier N, Lupascu A, Nista EC, Novi M, Candelli M, Cimica V, Mihm S, Gasbarrini G,<br />

Ramadori G, Gasbarrini A.<br />

Characterization <strong>of</strong> gene expression pr<strong>of</strong>ile in rat Kupffer cells stimulated with IFN-α or IFN-γ. Dig Liver Dis. 2006<br />

Aug;38(8):563-77. Epub 2006 Jun 27.<br />

Zwacka RM, Zhang Y, Halldorson J, Schlossberg H, Dudus L, Engelhardt JF.<br />

CD4 + T-lymphocytes mediate ischemia/reperfusion-induced inflammatory responses in mouse liver. J Clin Invest.<br />

1997 Jul 15;100(2):279-89.<br />

232


Maren Depke<br />

P U B L I C A T I O N S<br />

Peer-review articles<br />

Depke M, Fusch G, Domanska G, Geffers R, Völker U, Schuett C, Kiank C.<br />

Hypermetabolic syndrome as a consequence <strong>of</strong> repeated psychological stress in mice.<br />

Endocrinology. 2008 Jun;149(6):2714-23. Epub 2008 Mar 6. PMID: 18325986<br />

Depke M, Steil L, Domanska G, Völker U, Schütt C, Kiank C.<br />

Altered hepatic mRNA expression <strong>of</strong> immune response and apoptosis-associated genes after acute and<br />

chronic psychological stress in mice.<br />

Mol Immunol. 2009 Sep;46(15):3018-28. Epub 2009 Jul 9. PMID: 19592098<br />

Grabarczyk P, Przy<strong>by</strong>lski GK, Depke M, Völker U, Bahr J, Assmus K, Bröker BM, Walther R, Schmidt CA.<br />

Inhibition <strong>of</strong> BCL11B expression leads to apoptosis <strong>of</strong> malignant but not normal mature T cells.<br />

Oncogene. 2007 May 31;26(26):3797-810. Epub 2006 Dec 18. PMID: 17173069<br />

Grabarczyk P, Nähse V, Delin M, Przy<strong>by</strong>lski G, Depke M, Hildebrandt P, Völker U, Schmidt CA.<br />

Increased expression <strong>of</strong> Bcl11b leads to chemoresistance accompanied <strong>by</strong> G1 accumulation.<br />

PLoS One. 2010 Sep 2;5(9). pii: e12532.<br />

Depke M, Dinh Hoang Dang K, Breitbach K, Brinkmann L, Gesell Salazar M, Hammer E, Bast A, Steil L,<br />

Schmidt F, Steinmetz I, Völker U.<br />

Transcriptomic and proteomic <strong>characterization</strong> <strong>of</strong> mouse BMM after IFN-γ activation in a serum-free<br />

system.<br />

In preparation.<br />

Poster/Talk<br />

Analysis <strong>of</strong> the response <strong>of</strong> airway epithelial cells S9 to bacterial culture supernatants <strong>of</strong> Staphylococcus<br />

aureus RN1HG – impact <strong>of</strong> SigB<br />

M. Gutjahr, M. Depke, P. Hildebrandt, D. Hoppe, M. Degner, A. Kühn, E. Hammer, J. Pané-Farré, L. Steil,<br />

M. Hecker, U. Völker<br />

Meeting <strong>of</strong> Transregional Collaborative Research Center 34 “Pathophysiology <strong>of</strong> Staphylococci”<br />

Kloster Banz/Bad Staffelstein, Germany, 28.-31.10.2008 (Poster)<br />

Measuring expression signatures in <strong>host</strong>-<strong>pathogen</strong> <strong>interactions</strong><br />

M. Depke, U. Völker<br />

Summer School “Pathophysiologie von Staphylokokken in der Post-Genom-Ära”<br />

Vilm, Germany, 26.-29.09.2007 (Talk)<br />

Chronic stress-induced metabolic disturbances in female BALB/c mice<br />

M. Depke, C. Kiank, R. Geffers, C. Schütt, U. Völker<br />

2 nd International Alfried Krupp Kolleg Symposium “stress, behaviour, immune response“<br />

Greifswald, Germany, 9.-11.11.2006 (Poster)<br />

233


Maren Depke<br />

A F F I D A V I T / E R K L Ä R U N G<br />

Hiermit erkläre ich, daß diese Arbeit bisher von mir weder an der Mathematisch-<br />

Naturwissenschaftlichen Fakultät der Ernst-Moritz-Arndt-Universität Greifswald noch einer<br />

anderen wissenschaftlichen Einrichtung zum Zwecke der Promotion eingereicht wurde.<br />

Ferner erkläre ich, daß ich diese Arbeit selbständig verfaßt und keine anderen als die darin<br />

angegebenen Hilfsmittel benutzt habe.<br />

Greifswald, den 24. 09. 2010<br />

………………………………………………………………………………………………………<br />

Maren Depke<br />

235


Maren Depke<br />

A C K N O W L E D G M E N T S<br />

Many thanks go to all supervisors for the chance to work on this topic, for the mentoring and<br />

discussions, and notably for the possibility to use the excellent technical equipment in the<br />

Department <strong>of</strong> Functional Genomics. I like to thank all colleagues for their contribution to cooperation<br />

projects, and in particular for the nice working atmosphere, for the support and help in<br />

challenging times during this work. And last but not least, sincere thanks are given to all friends<br />

and especially to my family for the encouragement and care they gave to me.<br />

Interfaculty Institute for Genetics and Functional Genomics,<br />

Department <strong>of</strong> Functional Genomics, Ernst-Moritz-Arndt University <strong>of</strong> Greifswald, Germany<br />

Sabine Ameling<br />

Marc Burian<br />

Dinh Hoang Dang Khoa<br />

Melanie Gutjahr<br />

… and all other colleagues …<br />

Elke Hammer<br />

Petra Hildebrandt<br />

Ulrike Mäder<br />

Marc Schaffer<br />

Sandra Scharf<br />

Frank Schmidt<br />

Leif Steil<br />

Uwe Völker<br />

Liver gene expression pattern in a mouse psychological stress model<br />

Gene expression pr<strong>of</strong>iling was performed on mouse samples from a psychological stress<br />

model which was established and performed <strong>by</strong> Cornelia Kiank in the setting <strong>of</strong> the DFG-<br />

Graduiertenkolleg GK840 „Wechselwirkungen zwischen Erreger und Wirt bei generalisierten<br />

bakteriellen Infektionen“. After Affymetrix expression pr<strong>of</strong>iling method training <strong>by</strong> Tanja Töpfer<br />

and Robert Geffers, array hybridizations were conducted in the lab <strong>of</strong> the array facility at the<br />

Helmholtz-Zentrum für Infektionsforschung. Markus Grube provided the protocol for real-time<br />

PCR mastermix. The work was supervised <strong>by</strong> Christine Schütt, Barbara Bröker, and Uwe Völker.<br />

Cornelia Kiank, Christine Schütt, Barbara Bröker (Institute for Immunology and Transfusion Medicine,<br />

Department <strong>of</strong> Immunology, Ernst-Moritz-Arndt University <strong>of</strong> Greifswald, Germany; C. K. now: David<br />

Geffen School <strong>of</strong> Medicine, UCLA; Digestive Diseases Research Center and Center for Neurobiology <strong>of</strong><br />

Stress; Los Angeles, CA, USA)<br />

Tanja Töpfer, Robert Geffers, Jan Buer, Jürgen Wehland (Helmholtz Centre for Infection Research,<br />

Braunschweig, Germany)<br />

Markus Grube, Heyo Krömer (Institute <strong>of</strong> Pharmacology, Ernst-Moritz-Arndt University <strong>of</strong> Greifswald,<br />

Germany)<br />

Uwe Völker (Interfaculty Institute for Genetics and Functional Genomics, Department <strong>of</strong> Functional<br />

Genomics, Ernst-Moritz-Arndt University <strong>of</strong> Greifswald, Germany)<br />

237


Maren Depke<br />

Acknowledgments<br />

Kidney gene expression pattern in an in vivo infection model<br />

The mouse infection model was established and performed in the setting <strong>of</strong> the DFG<br />

Sonderforschungsbereich SFB TR 34 “Pathophysiologie von Staphylokokken in der Post-Genom<br />

Ära” <strong>by</strong> Tina Schäfer and Knut Ohlsen, who also contributed to study conception and design.<br />

Marc Burian co-operated in the wet-lab-work for the second biological replicate <strong>of</strong> infected<br />

samples and in the final data analysis and interpretation. The strain S. aureus RN1HG was<br />

generated in the lab <strong>of</strong> Friedrich Götz and its isogenic sigB mutant <strong>by</strong> Jan Pané-Farré in the lab <strong>of</strong><br />

Michael Hecker. The work was supervised <strong>by</strong> Uwe Völker.<br />

Tina Schäfer, Knut Ohlsen (Institute <strong>of</strong> Molecular Infection Biology, Julius-Maximilians University <strong>of</strong><br />

Würzburg, Germany)<br />

Jan Pané-Farré, Susanne Engelmann, Michael Hecker (Institute <strong>of</strong> Microbiology, Ernst-Moritz-Arndt<br />

University <strong>of</strong> Greifswald, Germany)<br />

Friedrich Götz (Department <strong>of</strong> Microbial Genetics, Eberhards-Karls University <strong>of</strong> Tübingen, Germany)<br />

Marc Burian, Uwe Völker (Interfaculty Institute for Genetics and Functional Genomics, Department <strong>of</strong><br />

Functional Genomics, Ernst-Moritz-Arndt University <strong>of</strong> Greifswald, Germany)<br />

Gene expression pattern <strong>of</strong> bone-marrow derived macrophages after interferon-γ activation<br />

Bone-marrow derived macrophages were generated and stimulated with IFN-γ <strong>by</strong> Kathrin<br />

Breitbach. Proteome analysis was conducted <strong>by</strong> Dinh Hoang Dang Khoa. The study was<br />

performed in the setting <strong>of</strong> the DFG Sonderforschungsbereich SFB TR 34 “Pathophysiologie von<br />

Staphylokokken in der Post-Genom Ära”. The work was supervised <strong>by</strong> Uwe Völker.<br />

Kathrin Breitbach, Antje Bast, Ivo Steinmetz (Institute <strong>of</strong> Medical Microbiology, Ernst-Moritz-Arndt<br />

University <strong>of</strong> Greifswald, Germany)<br />

Dinh Hoang Dang Khoa, Lars Brinkmann, Manuela Gesell Salazar, Elke Hammer, Leif Steil, Frank<br />

Schmidt, Uwe Völker (Interfaculty Institute for Genetics and Functional Genomics, Department <strong>of</strong><br />

Functional Genomics, Ernst-Moritz-Arndt University <strong>of</strong> Greifswald, Germany)<br />

Host cell gene expression pattern in an in vitro infection model<br />

Infection experiments were performed in close collaboration with Melanie Gutjahr, who<br />

worked on <strong>host</strong> proteome analysis and additionally accomplished control measurements and cell<br />

culture. Petra Hildebrandt conducted FACS cell sorting. The strain S. aureus RN1HG was<br />

generated in the lab <strong>of</strong> Friedrich Götz and its GFP expressing variant RN1HG pMV158GFP <strong>by</strong><br />

Leonard Menschner in the lab <strong>of</strong> Michael Hecker. The work was supervised <strong>by</strong> Uwe Völker. This<br />

study was performed in the setting <strong>of</strong> the DFG Sonderforschungsbereich SFB TR 34<br />

“Pathophysiologie von Staphylokokken in der Post-Genom Ära”.<br />

Leonhard Menschner, Susanne Engelmann, Michael Hecker (Institute <strong>of</strong> Microbiology, Ernst-Moritz-<br />

Arndt University <strong>of</strong> Greifswald, Germany)<br />

Melanie Gutjahr, Petra Hildebrandt, Elke Hammer, Uwe Völker (Interfaculty Institute for Genetics and<br />

Functional Genomics, Department <strong>of</strong> Functional Genomics, Ernst-Moritz-Arndt University <strong>of</strong> Greifswald,<br />

Germany)<br />

238


Maren Depke<br />

Acknowledgments<br />

Pathogen gene expression pr<strong>of</strong>iling – growth media and in vitro infection experiment studies<br />

This study was performed in the setting <strong>of</strong> the DFG Sonderforschungsbereich SFB TR 34<br />

“Pathophysiologie von Staphylokokken in der Post-Genom Ära” and the international<br />

cooperation EU-IP-FP6-project BaSysBio (LSHG-CT2006-037469).<br />

Infection experiments were performed in close collaboration with Melanie Gutjahr, who<br />

additionally accomplished control measurements, and Petra Hildebrandt, who conducted<br />

eukaryotic cell culture. Many thanks are given to Marc Schaffer for his valuable help during<br />

preparation <strong>of</strong> internalized staphylococci, to Ulrike Mäder for her help concerning all aspects <strong>of</strong><br />

tiling array data, and to Marc Burian for the discussions on staphylococcal gene expression. The<br />

work was supervised <strong>by</strong> Uwe Völker.<br />

Jan Pané-Farré, Susanne Engelmann, Michael Hecker (Institute <strong>of</strong> Microbiology, Ernst-Moritz-Arndt<br />

University <strong>of</strong> Greifswald, Germany)<br />

Magda M. van der Kooi-Pol, Annette Dreisbach, Jan Maarten van Dijl (Department <strong>of</strong> Medical<br />

Microbiology, University Medical Center Groningen / UMCG, Groningen, The Netherlands; A. D.<br />

formerly: Department <strong>of</strong> Functional Genomics, Ernst-Moritz-Arndt University <strong>of</strong> Greifswald, Germany)<br />

Hanne Jarmer (Center for Biological Sequence Analysis, Department <strong>of</strong> Systems Biology, Technical<br />

University <strong>of</strong> Denmark, Lyng<strong>by</strong>, Denmark)<br />

Pierre Nicolas, Aurélie Leduc, Philippe Bessières (INRA, Mathématique Informatique et Génome, Jouyen-Josas,<br />

France)<br />

Melanie Gutjahr, Petra Hildebrandt, Marc Schaffer, Marc Burian, Ulrike Mäder, Uwe Völker (Interfaculty<br />

Institute for Genetics and Functional Genomics, Department <strong>of</strong> Functional Genomics, Ernst-Moritz-<br />

Arndt University <strong>of</strong> Greifswald, Germany)<br />

FURTHER COOPERATION PARTNERS<br />

Piotr Grabarczyk, Christian A. Schmidt (Molecular Hematology, Department <strong>of</strong> Hematology and<br />

Oncology, Ernst-Moritz-Arndt University <strong>of</strong> Greifswald, Germany)<br />

FINANCIAL SUPPORT<br />

DFG-Graduiertenkolleg / Research Training Group GK840<br />

„Wechselwirkungen zwischen Erreger und Wirt bei generalisierten bakteriellen Infektionen“/<br />

“Host-<strong>pathogen</strong> <strong>interactions</strong> in generalized bacterial infection“<br />

DFG Sonderforschungsbereich / Transregional Collaborative Research Center SFB TR 34<br />

“Pathophysiologie von Staphylokokken in der Post-Genom Ära”/<br />

„Pathophysiology <strong>of</strong> Staphylococci in the Post-Genomic Era“<br />

239

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!