thesis mtf 1 - Duke University

204
Protein S-Nitrosylation: Methods of Detection and Cellular Regulation by Michael Tcheupdjian Forrester Department of Biochemistry Duke University Date:_______________________ Approved: ___________________________ Jonathan S. Stamler, MD (Supervisor) ___________________________ Irwin Fridovich, PhD ___________________________ K. V. Rajagopalan, PhD ___________________________ Dennis J. Thiele, PhD ___________________________ Eric J. Toone, PhD Dissertation submitted in partial fulfillment of the requirements for the degree of doctor of philosophy in the Department of Biochemistry in the Graduate School of Duke University 2009

Transcript of thesis mtf 1 - Duke University

Page 1: thesis mtf 1 - Duke University

i

v

Protein S-Nitrosylation: Methods of Detection and Cellular Regulation

by

Michael Tcheupdjian Forrester

Department of Biochemistry Duke University

Date:_______________________ Approved:

___________________________ Jonathan S. Stamler, MD

(Supervisor)

___________________________ Irwin Fridovich, PhD

___________________________ K. V. Rajagopalan, PhD

___________________________ Dennis J. Thiele, PhD

___________________________ Eric J. Toone, PhD

Dissertation submitted in partial fulfillment of the requirements for the degree of doctor of philosophy in the Department of

Biochemistry in the Graduate School of Duke University

2009

Page 2: thesis mtf 1 - Duke University

i

v

ABSTRACT

Protein S-Nitrosylation: Methods of Detection and Cellular Regulation

by

Michael Tcheupdjian Forrester

Department of Biochemistry Duke University

Date:_______________________ Approved:

___________________________ Jonathan S. Stamler, MD

(Supervisor)

___________________________ Irwin Fridovich, PhD

___________________________ K. V. Rajagopalan, PhD

___________________________ Dennis J. Thiele, PhD

___________________________ Eric J. Toone, PhD

An abstract of a dissertation submitted in partial fulfillment of the requirements for the degree of doctor of philosophy in the Department of

Biochemistry in the Graduate School of Duke University

2009

Page 3: thesis mtf 1 - Duke University

Copyright by Michael T. Forrester

2009

Page 4: thesis mtf 1 - Duke University

iv

Abstract

Protein S-nitrosylation—the post-translational modification of cysteine thiols into

S-nitrosothiols—is a principle mechanism of nitric oxide-based signaling. Studies have

demonstrated myriad roles for S-nitrosylation in organisms from bacteria to humans, and

recent efforts have begun to elucidate how this redox-based modification is regulated

during physiological and pathophysiological conditions. This doctoral thesis is focused

on the 1) analysis of existing methodologies for the detection of protein S-nitrosylation;

2) development of new methodologies for the detection of protein S-nitrosylation and 3)

discovery of novel enzymatic mechanisms by which S-nitrosylation is regulated in vivo.

The specificity of the biotin switch technique, the mainstay assay for detecting S-

nitrosylation, was rigorously assessed and validated. This study was paramount as a

response to several published (though poorly grounded) criticisms of the biotin switch

technique. Separately presented is a unique resin-based assay for proteomic analysis of S-

nitrosylation (dubbed “SNO-RAC”), which is combined with mass spectrometric tools to

identify sites of S-nitrosylation in several cellular models (e.g. E. coli, mammalian cells).

Other data presented herein demonstrate that the thioredoxin system is regulated, in a

negative feedback manner, to control S-nitrosylation and prevent nitrosative stress. This

system involves nitric oxide-dependent suppression of an established thioredoxin

inhibitor, the thioredoxin interacting protein (Txnip). This, in turn, affords thioredoxin an

optimal environment to drive protein denitrosylation and prevent nitrosative stress

secondary to endogenous nitric oxide production.

Page 5: thesis mtf 1 - Duke University

v

Contents Abstract .............................................................................................................................. iv List of Tables ................................................................................................................... viii List of Figures .................................................................................................................... ix List of Abbreviations ........................................................................................................ xii Acknowledgements.......................................................................................................... xiv

1. Introduction to S-Nitrosylation ....................................................................................... 1

1.1 Cysteine Reactivity and Redox Biochemistry ......................................................... 1 1.2 Nitric Oxide Synthases and S-Nitrosylation ............................................................ 2 1.3 Redox Stress as a Microbicidal Host Defense Mechanism...................................... 3 1.4 Microbial Strategies to Counteract Host-Derived ROS/RNS .................................. 5 1.5 Protein Denitrosylation and Nitrosative Stress Tolerance ....................................... 6

2. Assessment and Validation of the Biotin Switch Technique........................................ 12

2.1 Summary ................................................................................................................ 13 2.2 Introduction ............................................................................................................ 14 2.3 Criticisms Regarding the Specificity of the BST................................................... 15 2.4 Materials and Methods........................................................................................... 16 2.5 Specificity of the BST for the Detection of Protein SNOs .................................... 19 2.6 Conditions that Promote SNO-Independent Biotinylation .................................... 21 2.7 Pre-Photolysis as a Control within the BST........................................................... 22 2.8 Effects of Oxidative Stresses on Protein S-Nitrosylation ...................................... 22 2.9 Discussion .............................................................................................................. 24

3. Sulfur-Based SNO Detection: The Biotin Switch Technique....................................... 34

3.1 Summary ................................................................................................................ 35 3.2 NO- vs. Sulfur-Based Assays of S-Nitrosylation................................................... 36 3.3 Workflow of the Biotin Switch Technique ............................................................ 38 3.4 Pre-photolysis as an Independent Strategy in the BST .......................................... 39 3.5 Design of the Pre-photolysis Apparatus................................................................. 41 3.6 Methodological Concerns Raised about the BST .................................................. 42 3.7 Potential Problems with Exogenous Metals in the BST ........................................ 46

Page 6: thesis mtf 1 - Duke University

vi

3.8 Modern Permutations and Adaptations of the BST ............................................... 47 3.9 Sample Protocol for the BST ................................................................................. 48

4. Proteomic Analysis of S-Nitrosylation by Resin-Assisted Capture.............................. 59

4.1 Summary ................................................................................................................ 60 4.2 Introduction ............................................................................................................ 61 4.3 Materials and Methods........................................................................................... 62 4.4 An In Vitro Analysis of SNO-RAC with Purified GAPDH .................................. 67 4.5 Assessment of Intracellular S-Nitrosylation by SNO-RAC................................... 67 4.6 Assessment of Endogenous S-Nitrosylation by SNO-RAC................................... 68 4.7 Detection of S-Acylation by a Modified SNO-RAC: “Acyl-RAC” ...................... 70 4.8 Mapping SNO-Sites by SNO-RAC and LC-MS/MS............................................. 70 4.9 Combining SNO-RAC with Isotopically Encoded LC-MS/MS ............................ 72 4.10 Discussion ............................................................................................................ 73

5. Regulated Protein Denitrosylation by the Thioredoxin System ................................... 90

5.1 Summary ................................................................................................................ 91 5.2 Introduction ............................................................................................................ 92 5.3 Materials and Methods........................................................................................... 92 5.4 Identification of Trx1 as a denitrosylating enzyme.............................................. 101 5.5 Trx1 mediates basal denitrosylation in intact cells .............................................. 103 5.6 Denitrosylation proceeds, at least in part, via an intermolecular disulfide .......... 105 5.7 Trx1 is a broad-spectrum denitrosylase ............................................................... 106 5.8 Trx2 mediates stimulus-induced denitrosylation of caspase-3 and apoptosis ..... 106 5.9 Discussion ............................................................................................................ 108

6. Txnip is a Feedback Regulator of S-Nitrosylation...................................................... 122

6.1 Summary .............................................................................................................. 123 6.2 Introduction .......................................................................................................... 124 6.3 Thioredoxin is an Enzymatic Denitrosylase ........................................................ 125 6.4 Nitric Oxide Synthesis and Nitrosative Stress ..................................................... 126 6.5 Thioredoxin Interacting Protein: An Introduction ............................................... 127 6.6 Materials and Methods......................................................................................... 127 6.7 Txnip Expression is Negatively Regulated by Endogenous NO ......................... 132 6.8 Txnip Suppression is Replicated by Exogenous NO ........................................... 134 6.9 Txnip Suppression is Independent of O2 and cGMP ........................................... 135 6.10 Txnip Suppression Involves mRNA Downregulation ....................................... 136

Page 7: thesis mtf 1 - Duke University

vii

6.11 Txnip Protein Undergoes Rapid Intracellular Turnover .................................... 136 6.12 Nitric Oxide Inhibits the Txnip Promoter .......................................................... 137 6.13 Txnip Repression Protects from Nitrosative Stress ........................................... 138 6.14 Txnip Binds to the Active Site Cys32 of Trx1.................................................... 139 6.15 Txnip Potentiates Intracellular Protein S-Nitrosylation ..................................... 140 6.16 Discussion .......................................................................................................... 141

References....................................................................................................................... 165

Biography........................................................................................................................ 189

Page 8: thesis mtf 1 - Duke University

viii

List of Tables

Table 1: Identification of high molecular weight protein-SNOs ...................................... 83

Table 2: Identification of SNO-sites in murine macrophages .......................................... 84

Table 3: Identification of SNO-sites in E. coli ................................................................. 87

Page 9: thesis mtf 1 - Duke University

ix

List of Figures

Figure 1: Cys-based redox modifications ........................................................................... 8

Figure 2: Basic mechanisms of protein S-nitrosylation ...................................................... 9

Figure 3: The reactions of S-nitrosoglutathione reductase (GSNOR) .............................. 10

Figure 4: Basic mechanisms of protein denitrosylation.................................................... 11

Figure 5: Specificity of the BST in vitro and in lysates.................................................... 29

Figure 6: Specificity of the BST in vivo ........................................................................... 30

Figure 7: Indirect sunlight drives “artifactual” biotinylation............................................ 31

Figure 8: Pre-photolysis eliminates SNO-dependent biotinylation .................................. 32

Figure 9: S-nitrosylation of GAPDH is promoted by oxidative stress.............................. 33

Figure 10: Sulfur- vs. NO-based assays for S-nitrosylation ............................................. 54

Figure 11: Overview of the biotin switch technique (BST).............................................. 55

Figure 12: Transnitrosation between SNOs and ascorbate ............................................... 56

Figure 13: Pre-photolysis and ascorbate controls within the BST.................................... 57

Figure 14: Results from a typical BST ............................................................................. 58

Figure 15: Schematic illustration of SNO-RAC ............................................................... 76

Figure 16: SNO-RAC maps SNO-sites in vitro ................................................................ 77

Figure 17: A side-by-side comparison of the BST and SNO-RAC.................................. 78

Figure 18: SNO-RAC detects endogenous protein-SNOs................................................ 79

Page 10: thesis mtf 1 - Duke University

x

Figure 19: Detection of S-acylated proteins by Acyl-RAC .............................................. 80

Figure 20: SNO-RAC combined with label-free LC-MS/MS .......................................... 81

Figure 21: SNO-RAC combined with isotopic MALDI MS............................................ 82

Figure 22: Identification of Trx1 as a caspase-3 denitrosylase....................................... 109

Figure 23: Characterization of caspase-3 denitrosylase activity..................................... 110

Figure 24: Partial purification and enrichment of caspase-3 denitrosylating activity .... 111

Figure 25: Proteins in “Fraction I” identified by MALDI-TOF MS .............................. 112

Figure 26: Modulation of Trx1 activity determines caspase-3 denitrosylation .............. 113

Figure 27: The active site of Trx1 is critical for denitrosylation .................................... 114

Figure 28: Trx1-TrxR1 mediates protein denitrosylation in intact cells......................... 115

Figure 29: Trx-TrxR inhibition increases caspase-3-SNO in vivo ................................. 116

Figure 30: Interaction of Trx1 with SNO-proteins ......................................................... 117

Figure 31: Trx-TrxR inhibition increases the multiple proteins-SNOs in vivo .............. 118

Figure 32: Trx2 mediates Fas-induced denitrosylation of mitochondria caspase-3 ....... 119

Figure 33: TrxR2 inhibition blocks Fas-induced denitrosylation of caspase-3 .............. 120

Figure 34: TrxR2 mediates caspase activation and DNA fragmentation ....................... 121

Figure 35: Disulfide reduction by the thioredoxin system.............................................. 122

Figure 36: Protein denitrosylation by the thioredoxin system........................................ 123

Figure 37: Thioredoxin mediates protein denitrosylation in cells .................................. 124

Page 11: thesis mtf 1 - Duke University

xi

Figure 38: Thioredoxin-interacting protein (Txnip) in macrophages ............................. 147

Figure 39: Txnip is suppressed in HEK293 cells............................................................ 148

Figure 40: A time-course study in stimulated murine macrophages .............................. 149

Figure 41: Macrophages from iNOS-/- mice fail to suppress Txnip ............................... 150

Figure 42: Txnip suppression is replicated by exogenous NO ....................................... 151

Figure 43: The role of O2 in Txnip suppression ............................................................. 152

Figure 44: The role of cGMP in Txnip suppression ....................................................... 153

Figure 45: RT-PCR of Txnip mRNA in response to NO synthesis................................ 154

Figure 46: A time-course study of Txnip protein stability ............................................. 155

Figure 47: NO inhibits Txnip promoter activity ............................................................. 156

Figure 48: Txnip facilitates LDH release in response to iNOS ...................................... 157

Figure 49: Txnip does not potentiate iNOS activity ....................................................... 158

Figure 50: Txnip inhibits DNA synthesis in response to iNOS...................................... 159

Figure 51: Cellular morphology during Txnip-driven nitrosative stress ........................ 160

Figure 52: Txnip binds to the active site Cys32 of Trx1.................................................. 161

Figure 53: Txnip potentiates general protein S-nitrosylation ......................................... 162

Figure 54: Txnip potentiates the S-nitrosylation of GAPDH.......................................... 163

Figure 55: A schematic overview of Txnip function ...................................................... 164

Page 12: thesis mtf 1 - Duke University

xii

List of Abbreviations: BCNU – 1,3-bis(2-chloroethyl)-1-nitrosourea

biotin-HPDP – N-[6-(biotinamido)hexyl]-3'-(2'-pyridyldithio)propionamide

8-bromo-cGMP – 8-Bromoguanosine-3',5'-cyclic monophosphate

BSA – bovine serum albumin

BST – biotin switch technique

CysNO – S-nitrosocysteine

DETA-NO – diethylenetriamine NONOate

DTNB – 5,5’-dithiobis(2-nitrobenzoic acid)

DTT – dithiothreitol

eNOS – endothelial nitric oxide synthase

GAPDH – glyceraldehyde-3-phosphate dehydrogenase

GSH – reduced glutathione

GSSG – glutathione disulfide

IFN-γ – interferon-γ

iNOS – inducible nitric oxide synthase

LPS – lipopolysaccharide

M-CSF – macrophage colony stimulating factor

MMTS – S-methyl methanethiosulfonate

NADH – nicotinamide adenine dinucleotide, reduced

NADPH – nicotinamide adenine dinucleotide phosphate, reduced

Page 13: thesis mtf 1 - Duke University

xiii

nNOS – neuronal nitric oxide synthase

NO – nitric oxide

ODQ – 1H-[1,2,4]oxadiazolo[4,3-a]quinoxalin-1-one

pNPP – p-nitrophenyl phosphate

PTP1B – human protein tyrosine phosphatase 1B

SNO – S-nitrosothiol

Trx – thioredoxin

TrxR – thioredoxin reductase

Txnip – thioredoxin-interacting protein

Page 14: thesis mtf 1 - Duke University

xiv

Acknowledgements

Training as an MD/PhD student has been a tremendous gift. This incredible

experience would never have been granted to me without the unwavering support and

guidance of many wonderful individuals, and without them I would be nowhere.

To my parents (Leon and Linda) and my sister (Leslie), I am forever indebted for

providing everything an individual could ever need (especially good genes!). Education

was never forced upon me as a youth, and I believe it is for this reason that I learned to

truly embrace education while most of my contemporaries became either exhausted or

disgusted by it. My father also taught me how to be unshakable.

I am very grateful to the former director of the Medical Scientist Training

Program (MSTP), Salvatore Pizzo, for accepting me into his training program and being

a constant source of inspiration during my early (and most daunting) days at Duke. I also

thank the entire Department of Biochemistry for welcoming me into their academic home

and providing me with every possible opportunity for success. Just as well, my

committee has been very supportive and genuinely interested in helping me become a

scientist.

Within Dr. Stamler’s laboratory, I am particularly grateful to several individuals

for taking me “under their wing” and sharing their wisdom. When I arrived in the lab, I

had almost no experience in biochemistry or cellular biology. Matthew Foster, a research

associate in the lab, literally taught me how to perform experiments…and most

importantly, how to design controls! Douglas Hess, a senior research associate in the lab,

Page 15: thesis mtf 1 - Duke University

xv

has been instrumental in my learning how to critically interpret and present experimental

data. Another research associate, Moran Benhar, has been an excellent scientific

companion from whom I have learned a great deal and have been able to contribute to

several exciting projects. I am also indebted to many other members of the laboratory for

their constant support and advice. They include (in no particular order): Divya Seth,

Alfred Hausladen, Laura Dillon, Michael Angelo, Balakrishnan Selvakumar, Akio

Matsumoto, Kentaro Ozawa, Liang Xie, Qi-An Sun, Zhiqiang Chen, Harvey Marshall,

Naoko Adachi, Bo Li, Puneet Anand, Ben Chung, Abhijit Chakladar, Diana Diesen, Kyla

Bennett, Sung Oog Kim, and Leonardo Nogueira. My classmates Jane Healy and

Christine Eyler have also been very helpful in providing experimental expertise and

motivation.

It has been a true pleasure to study under my supervisor, Jonathan Stamler. He has

provided me with myriad opportunities that taught me how to function as a scientist. Dr.

Stamler has continuously offered me ample support and guidance, yet he always gave me

space to work out my own ideas and become proficient in functioning independently. But

above all other things, he has motivated me to forget how to be a follower (which

frequently becomes engrained during undergraduate and medical training) and to never

be afraid of pursuing the questions that others avoid. For me these are lifelong lessons,

and I will carry them with me regardless of professional goals or academic titles.

Page 16: thesis mtf 1 - Duke University

1

Chapter 1. Introduction to S-Nitrosylation

1.1 Cysteine Reactivity and Redox Biochemistry

Cysteine (Cys) is a highly unique amino acid due to its thiol side chain. This

functional group is nucleophilic, acidic (pKa ~ 8) and redox active due to its hybridized

p- and d-orbitals, which together impart a vast range of reactivities for Cys residues

within proteins. Importantly, these properties of Cys residues can be affected by protein

secondary, tertiary or quaternary structure. For example, numerous proteins employing an

active site Cys residue lower its pKa by several orders of magnitude (to 5 or 6) via

tertiary interactions with basic groups that promote thiol deprotonation. These include

Cys-hydrolases such as caspases and calpains, as well as redox-reactive proteins such as

thioredoxin and glutaredoxin (to be discussed later).

Within the realm of redox-chemistry (i.e., transfer of electrons and consequent

change in atomic oxidation state), numerous reactions are known to occur on Cys thiol

side chains, which may affect protein structure, activity and function. The biologically

relevant Cys modifications are depicted in Figure 1. Illustrated are oxidative

modifications such as disulfide formation (e.g., glutathionylation) and S-oxides such as

sulfenic, sulfinic and sulfonic acids. These modifications are typically driven by reactive

oxygen species (ROS) like superoxide (O2-), hydrogen peroxide (H2O2) or hydroxyl

radical (•OH). These ROS are constitutively generated as a consequence of aerobic

metabolism (i.e., reduction of O2), as well as during conditions of oxidative stress or

infection (e.g., via the NADPH-oxidase system) [1, 2]. Though the generation of ROS

certainly bears biological consequences, the utility of ROS as signaling modalities is less

well established.

Page 17: thesis mtf 1 - Duke University

2

1.2 Nitric Oxide Synthases and S-Nitrosylation

Illustrated in Figure 1 is an S-nitrosothiol (SNO), which is the adduct of a nitroso

moiety and a Cys thiolate. Relative to the other oxidative modifications listed in Figure 1,

an SNO is unique in several major respects. Firstly and most obviously, it contains a

nitrogen atom. Secondly, the SNO bond is resonance delocalized via two major

resonance structures (S-N=O and +S=N-O-), which imparts some degree of stability to the

SNO moiety. Thirdly, the S-NO bond can undergo both homolytic and heterolytic

cleavage reactions (discussed below). Lastly, an SNO is fully reversible back to thiol via

a unique type of transnitrosation reaction (discussed below).

The conversion of a Cys thiol (or thiolate) to an SNO can be achieved by several

reactions. In the laboratory, low molecular weight thiols (e.g. Cys, GSH) are readily S-

nitrosated via mixing with nitrite (NO2-) under acidic conditions, which generates potent

nitrosants such as nitrosonium ion (NO+) and dinitrogen trioxide (N2O3). This reaction,

however, requires very low pH (i.e. < 3) and is therefore likely not physiologically

relevant, with the major exception of the stomach. Another route to generate an SNO

involves a reaction between nitric oxide (NO) and a Cys thiol where the redox

requirement is met by loss of one electron (Figure 2A). The three nitric oxide synthase

isoforms (i.e. nNOS/NOS1, iNOS/NOS2, eNOS/NOS3) are all capable of driving S-

nitrosylation in vivo and in vitro, which is direct evidence that NO can undergo redox

reactions necessary to form SNOs. The three NOS isoforms are heme- and flavin-

containing proteins that employ NAPDH, tetrahydrobiopterin and O2 to convert L-

arginine to L-citrulline with concomitant release of NO [3]. Once formed, an SNO may

readily react with another thiol via a transnitrosation reaction (Figure 2B). This reaction

Page 18: thesis mtf 1 - Duke University

3

forms the basis by which SNO moieties may be rapidly transferred between between

proteins or low molecular weight thiols.

1.3 Redox Stress as a Microbicidal Host Defense Mechanism

A major component of mammalian innate immunity involves the enzymatic

synthesis of ROS and reactive nitrogen species (RNS), prototypic of which are O2-/H2O2

and NO/SNO, respectively. The production of ROS by mammalian phagocytes, dubbed

the “respiratory burst” due to massively elevated O2 consumption/reduction, allows

neutrophils and macrophages to kill invading microorganisms. Humans deficient in the

ROS-generating NADPH-oxidase (NOX) system suffer from chronic granulomatous

disease, a genetic condition that predisposes to recurrent bacterial infections [4, 5]. In

contrast to ROS, the synthesis of RNS is widespread from epithelium to lymphocytes.

Inducible nitric oxide synthase (iNOS), the enzyme that produces high-output NO (i.e.

nano- to micromolar), is synthesized in many cell types following diverse stresses such as

infection [6, 7], ionizing radiation [8, 9], liver resection [10] and neurologic trauma [11].

Further, humans that express lower levels of iNOS, due to polymorphisms in the iNOS

promoter region, are hypersensitive to malaria [12-15]. In each of these cases, iNOS

appears to play a protective role, though a multitude of other studies have suggested that

dysregulated or excessive iNOS activity may exert deleterious effects as well.

There exist a wide range of chemical mechanisms by which ROS and RNS might

impose a redox stress. These may be grouped by the nature of the target (e.g. protein,

DNA, lipid), the nature of the modification (e.g. S-nitroso, N-nitroso, S-oxy, N-oxy, C-

oxy, C-nitro, disulfide) or the specific chemical intermediate responsible for damage (e.g.

Page 19: thesis mtf 1 - Duke University

4

H2O2 from NOX, •OH from Fenton reactions, N2O3 from NO autooxidation, ONOO-

from NO and O2-). For example, H2O2 is known to oxidize Cys thiols to disulfides and S-

oxides such as sulfenic (-SOH), sulfinic (-SO2H) and sulfonic (-SO3H) acids. H2O2 can

also drive cell death by Fe2+-dependent Fenton reactions leading to •OH (or perferryl

radical, •OOH) production, which in turn hydroxylates DNA bases or causes backbone

scission via abstraction of the 2’ hydrogen atoms [16, 17]. Importantly, NO and low

molecular weight SNOs such as S-nitrosocysteine (CysNO) appear to inhibit cellular

growth differently than ROS, though several studies have demonstrated that they too can

facilitate DNA damage [18], likely by promoting Fenton chemistry via inhibition of

oxidative respiration and accumulation of reducing equivalents (e.g. NADH) [19].

Reactive nitrogen species have also been shown to rapidly S-nitrosylate intracellular

protein Cys thiols and, in the case of free NO, to ligate both heme and non-heme Fe

atoms [20-23]. Theoretically, proteins with critical Cys residues could be affected by S-

nitrosylation, and enzymatic activities that require a reduced Cys thiol would be

inhibited, as occurs in mammalian caspases, for example [24-26]; another prime example

includes S-nitrosylation of GAPDH, which facilitates its nuclear localization and initiates

a cell death program [27, 28]. The Cys-based redox modifications described above, along

with their respective formal oxidation states, are illustrated in Figure 1.

A strong case has also been made for combinatorial reactions between ROS and

RNS, specifically the production of peroxynitrite (ONOO-). The radical species O2- and

NO readily react to form ONOO-, which is highly cytotoxic to bacteria [29] and induces a

Page 20: thesis mtf 1 - Duke University

5

wide range of chemical modifications; in particular, tyrosine nitration has received a

great deal of attention as potentially causative for ONOO- dependent death [30, 31].

1.4 Microbial Strategies to Counteract Host-Derived ROS/RNS

Microbes have evolved highly complex and elegant mechanisms to destroy ROS

and RNS, which allow their continued survival in the presence of oxidative and/or

nitrosative stresses. Superoxide dismutases (SODs), catalases (H2O2 dismutases) and

NAD(P)H-peroxidases directly catabolize O2-/H2O2, and these are critical ROS-protective

factors in nearly every known organism/pathogen [32-35]. Disruption of these protective

enzymes often sensitizes to ROS, though frequently numerous mutations must be made

due to functional redundancy and/or compensation when only one system is removed

[17]. In contrast to ROS, enzymes that catabolize RNS were only discovered during the

past decade. For example, bacteria and fungi respond to RNS by readily expressing an

NADH-dependent, NO consuming flavohemoglobin that converts cytotoxic NO to either

NO3- or N2O under aerobic and anaerobic conditions, respectively [36-40]. We and others

have shown the critical role of flavohemoglobins in protecting from nitrosative stress in

E. coli [39], S. aureus [41], P. aeruginosa [42] and C. neoformans [43].

Flavohemoglobins are virulence factors and mitigate protein modifications by NO [43,

44]. It is therefore established that microbial flavohemoglobins consume NO, prevent

excessive S-nitrosylation and ultimately protect the pathogen from RNS-based insult.

Page 21: thesis mtf 1 - Duke University

6

1.5 Protein Denitrosylation and Nitrosative Stress Tolerance

Another major step in understanding how microbes survive nitrosative stress was

the discovery that a highly conserved GSH-dependent alcohol dehydrogenase (E. coli

gene adhC) is also an S-nitrosoglutathione reductase (GSNOR) that protects from

nitrosative stresses in a multitude of organisms [45-47]. Importantly, this is the first

example of an enzyme that directly acts on an SNO molecule: it specifically reduces the

N=O bond of GSNO to an N-hydroxysulfenamide (Figure 3), which then decomposes to

either GSH or GSSG. This enzyme, however, acts only on GSNO (protein-SNOs are not

substrates), and it indirectly affects protein S-nitrosylation by modulating the equilibrium

between protein-SNO and GSNO [47, 48]. A recent report indicates that carbonyl

reductase also possesses GSNOR activity [49], raising the idea that multiple enzymes

may modulate GSNO levels in vivo.

Another equally important issue is how organisms might directly denitrosylate

protein-SNOs to reverse nitrosative stress. Certainly GSH plays a major role in removing

protein SNO groups, likely via transnitrosylation to produce GSNO, which undergoes

reduction by GSNOR. However, the concept of enzymatic mechanisms to remove

protein-SNO groups is relatively new and largely unexplored. There exist several likely

chemical mechanisms by which an SNO group may be removed (Figure 3).

The identification of a protein S-denitrosylase would be an important discovery

for several reasons. Firstly, it would establish that protein-SNOs (and not just GSNO) are

recognized by enzymes that metabolize SNOs. Secondly, such enzymes could determine

which protein-SNOs are stable and which are rapidly degraded. Thirdly, an S-

Page 22: thesis mtf 1 - Duke University

7

denitrosylase might allow the microbe to survive host-imposed nitrosative stress by

preventing excessive S-nitrosylation of specific SNO-sensitive targets.

Page 23: thesis mtf 1 - Duke University

8

Figure 1. Various (patho)physiologically-relevant Cys-based redox modifications. Depicted is a single Cys residue and side chain in a protein backbone. The formal oxidation state of each modification is listed in parentheses. These may occur on both high and low molecular weight species (e.g. proteins, glutathione). Clockwise, the modifications include S-nitrosylation to an S-nitrosothiol, S-glutathionylation (disulfide), S-oxidation to sulfenic, sulfinic and sulfonic acids.

Page 24: thesis mtf 1 - Duke University

9

Figure 2. Basic mechanisms of S-nitrosylation. Depicted are two routes for the conversion of a Cys thiol to an S-nitrosothiol (SNO). (A) Nitric oxide can undergo a redox reaction generating a higher N-oxide (e.g. N2O3), which in turn may nitrosate a Cys thiol to generate an SNO. Alternatively, the redox requirement may be satisfied by first oxidizing the Cys thiol to a thiyl radical, followed by radical propotionation of the latter with NO. (B) The nitroso group of an SNO can be transferred to another Cys thiol via transnitrosation (nitroso transfer). This reaction allows SNO groups to rapidly transfer between proteins and/or low molecular weight thiols (e.g. GSH).

Page 25: thesis mtf 1 - Duke University

10

Figure 3. The reactions of S-nitrosoglutathione reductase (GSNOR). (A) The enzyme reduces GSNO to GSNHOH (glutathione hydroxysulfenamide), which can undergo another reduction or decompose to GSSG and hydroxylamine or ammonia. (B), The same enzyme was originally described as a dehydrogenase that converts S-hydroxymethylglutathione to S-formylglutathione. This pathway is believed to play a role in formaldehyde detoxification. The enzyme is highly specific for GSH-based substrates and thus lacks activity towards protein-SNOs.

Page 26: thesis mtf 1 - Duke University

11

Figure 4. Proposed chemical mechanisms of protein S-denitrosylation. The shaded sphere represents a protein. Clockwise from top left: 1) Nucleophilic attack of a thiol at the sulfur atom of the SNO leads to a disulfide and nitroxyl anion (thiolation); 2) Transnitrosylation allows the nitroso group to undergo a concerted transfer reaction between thiols; 3) Reduction of S-N bond to thiol and NO or nitroxyl; 4) Hydrolysis of the S-N bond to yield thiol and nitrite; 5) Homolysis of the S-N bond to yield a transient thiyl radical and NO. Any of these routes may be enzyme catalyzed, presumably via stabilization of transition states and/or intermediate species.

Page 27: thesis mtf 1 - Duke University

12

Chapter 2. Assessment and Validation of the Biotin Switch Technique

Upon entering the laboratory, several recent publications had criticized the

specificity of the biotin switch technique (BST), the mainstay assay for detecting protein-

SNOs. These authors extrapolated their data to challenge the validity of ascorbate-based

conversion of SNO to thiol, which is a cornerstone component of the BST. Unfortunately,

their data were both thermodynamically ungrounded and likely explained by artifacts or

misinterpretations of the raw data (as discussed herein).

This chapter is based on a publication in the Journal of Biological Chemistry that

sought to independently assess the specificity of the biotin switch technique with purified

proteins, lysates and cells. The data presented here were partly in response to the

previously published criticisms of the BST. Also presented is a novel methodology (UV

“Pre-photolysis”), which further strengthens the specificity and rigor of the biotin switch

technique.

Forrester, M.T., Foster, M.W. & Stamler, J.S. Assessment and application of the biotin switch technique for examining protein S-nitrosylation under conditions of pharmacologically induced oxidative stress. J Biol Chem 282, 13977-13983 (2007).

Page 28: thesis mtf 1 - Duke University

13

2.1 Summary

Protein S-nitrosylation has emerged as a principal mechanism by which nitric

oxide exerts biological effects. Among methods for studying protein S-nitrosylation, the

biotin switch technique (BST) has rapidly gained popularity due to the ease with which it

can detect individual S-nitrosylated (SNO) proteins in biological samples. The

identification of SNO sites by the BST relies on the ability of ascorbate to generate a thiol

from an S-nitrosothiol, but not from alternatively S-oxidized thiols (e.g. disulfides,

sulfenic acids). However, the specificity of this reaction has recently been challenged,

prompting several claims that the BST may produce false-positive results and raising

concerns about the application of the BST under oxidizing conditions. Here we perform a

comparative analysis of the BST using differentially S-oxidized and S-nitrosylated forms

of protein tyrosine phosphatase 1B, as well as intact and lysed HEK-293 cells treated

with S-oxidizing and S-nitrosylating agents, and verify that the assay is highly specific for

SNO. Strikingly, exposure of samples to indirect sunlight from a laboratory window

resulted in artifactual ascorbate-dependent signals that are likely promoted by the

semidehydroascorbate radical; protection from sunlight eliminated the artifact. In

contrast, exposure of SNO-proteins to a strong ultraviolet light source (SNO-photolysis)

prior to the BST provided independent verification of assay specificity. By combining

BST with photolysis, we show that anticancer drug-induced oxidative stress facilitates the

S-nitrosylation of the major apoptotic effector GAPDH. Collectively, these experiments

demonstrate that SNO-dependent signaling pathways can be modulated by oxidative

conditions and suggests a potential role for S-nitrosylation in antineoplastic drug action.

Page 29: thesis mtf 1 - Duke University

14

2.2 Introduction

Nitric oxide (NO) executes a diverse range of cellular functions through the

redox-dependent conversion of protein Cys thiols to S-nitrosothiols (SNOs). This post-

translational modification, known as S-nitrosylation, has emerged as a highly conserved

and spatiotemporally specific signaling mechanism [21]. A major contribution to this

field was the introduction of the biotin switch technique (BST) by Jaffrey et al. in 2001

[50]. Importantly, the BST allows relatively facile identification and quantification of

endogenous protein SNOs, and thus has greatly contributed to our understanding of

protein S-nitrosylation. As evidence, the BST has been employed in many publications,

unveiling new roles for S-nitrosylation in events such as apoptosis [27],

neurodegeneration [51, 52], insulin signaling [53, 54] and receptor trafficking [55].

The BST consists of three principal steps (see Figure 11): “blocking” of free Cys

thiols by S-methylthiolation with MMTS (a reactive thiosulfonate), formal reduction of

SNOs to thiols with ascorbate, and in situ “labeling” by S-biotinylation of the nascent

thiols with biotin-HPDP, a reactive mixed disulfide. The degree of biotinylation (and

hence S-nitrosylation) is determined by either anti-biotin immunoblotting or streptavidin

pull-down followed by immunoblotting for the protein(s) of interest. Complete blocking

of free thiols is requisite for minimizing background biotinylation (i.e. improving signal-

to-noise or sensitivity), while the specificity of the BST is predicated on the ability of

ascorbate to reduce SNOs to free thiol, without reducing other Cys-based redox

modifications such as S-glutathionylation or S-oxidations (sulfenic, sulfinic and sulfonic

acids). Importantly, transition metal-catalyzed reactions, particularly those that might

Page 30: thesis mtf 1 - Duke University

15

limit the specificity of ascorbate, are mitigated through use of metal chelators (e.g.

EDTA, DTPA, neocuproine).

2.3 Criticisms Regarding the Specificity of the BST

Despite the unequivocal ability of the BST to detect SNOs, the specificity of

ascorbate for SNO reduction has not been extensively investigated. Landino et al.

suggested that ascorbate reduces tubulin disulfides, a claim the authors use to challenge

the validity of the BST [56]. Furthermore, two other groups have reported that S-

glutathionylated proteins yield positive signals in this assay [57, 58]. A more recent

report by Huang et al. argued that the BST yields “artifactual” ascorbate-dependent

biotinylation of native reduced BSA [59]. They extrapolate this finding to argue that the

BST may result in false-positive (i.e. SNO-independent) biotinylation, which could cast

doubt upon the myriad of studies that have relied heavily on this assay.

Shared between these claims of “artifactual signals” is the idea that ascorbate

must be reducing disulfides (S-glutathionylated or intramolecular disulfides). However,

this notion is difficult to reconcile experimentally with the use of MMTS as a blocking

agent (i.e. protein thiols are S-methylthiolated to form mixed disulfides). The assay

would never work if ascorbate removed the blocking agent, and in fact, the two e-

reduction of Cys disulfides to thiol (Cys thiol/disulfide Ered -170 to -320 mV) by

ascorbate is highly thermodynamically unfavorable (ascorbate/dehydroascorbate Ered +70

mV [60]). Thus, one would not expect ascorbate to directly reduce any biological Cys

oxidation products. Instead, the electrochemical measurements favor the reverse reaction

Page 31: thesis mtf 1 - Duke University

16

(i.e. thiol-dependent reduction of dehydroascorbate to ascorbate)—a scenario supported

by extensive observations [61, 62].

Despite the thermodynamic and empiric evidence against ascorbate-dependent

disulfide reduction (and alternative interpretations for some of the prior results that have

formed the basis of challenge to the BST), we felt it was important to evaluate the

specificity of the BST and the sources of potential artifacts. Further, we sought to employ

the BST to differentiate S-nitrosylated from S-oxidized proteins under conditions of

diverse oxidative and nitrosative stresses, as such analyses have not been previously

possible.

2.4 Materials and Methods

Materials – All materials were from Sigma unless otherwise indicated. Anti-

biotin, anti-pentaHis, anti-GSH, anti-PTP1B and anti-GAPDH monoclonal antibodies

were from Vector Laboratories, Qiagen, Virogen, BD Biosciences, and Chemicon,

respectively. Ni-NTA resin was from Qiagen. The cDNA sequence encoding the N-

terminal 321 residues of PTP1B in pET14b was a kind gift from S. G. Rhee (NIH).

CysNO was freshly synthesized by mixing equimolar volumes of L-Cys in 0.5 M HCl

and NaNO2. BCNU was freshly prepared as a 50 mM stock in EtOH and used

immediately.

Cell Culture and Treatments – Wild-type or stably nNOS-expressing [63] HEK-

293 cells were cultured in Dulbecco’s Modified Essential Media supplemented with 10%

FBS and treated with indicated agents at ~90% confluency. Cells were washed and

Page 32: thesis mtf 1 - Duke University

17

collected with cold PBS followed by lysis in HEK lysis buffer (25 mM HEPES, 50 mM

NaCl, 0.1 mM EDTA, 0.1 mM neocuproine, 0.1% NP-40, 1 mM PMSF).

Purification and Characterization of PTP1B – A 1L culture of BL21 E. coli

harboring pET14b-PTP1B at OD600 = 1.0 was induced with 1 mM IPTG for 4 h at 37 ºC

followed by purification of PTP1B with Ni-NTA resin as described by Qiagen. The

purified protein (4.5 mg/ml) was treated with 10 mM DTT for 30 min followed by buffer

exchange with Bio-Gel P-6DG (Bio-Rad) equilibrated in HEN buffer (250 mM HEPES,

1 mM EDTA, 0.1 mM neocuproine, pH 7.7). For measurement of phosphatase activity,

PTP1B was diluted to 2.5 ug/ml with phosphatase assay buffer (100 mM Tris, 2 mM

EDTA, 5 mM pNPP, pH 8.0) and optical absorbance was measured continuously at 405

nm. PTP1B S-glutathionylation was assessed by immunoblotting with anti-GSH antibody

(1:1000). PTP1B S-nitrosylation was confirmed using the Saville assay. Samples were

mixed with one vol. of 1% sulfanilamide with 2 mM HgCl2 in 1 M HCl for 5 min

followed by addition of one vol. of 0.05% N-(1-napthyl)ethylenediamine in 0.5 M HCl

and absorbance was read at 540 nm.

Protein Modifications and Biotin Switch Technique – Proteins/lysates were

treated with the indicated concentrations of H2O2, diamide, GSSG or CysNO followed by

desalting with Bio-Gel P-6DG. The BST was performed on purified PTP1B as described

[64] and with several modifications as follows.

Samples were mixed with 2 vol. of HEN buffer followed by addition of freshly

prepared MMTS (10% v/v in DMF) and SDS (25% v/v) to final concentrations of 0.1%

and 2.5%, respectively. Following frequent vortexing at 50 0C for 20 min, proteins were

precipitated with 3 vol. acetone at -20 0C for 15 min. The proteins were recovered by

Page 33: thesis mtf 1 - Duke University

18

centrifugation at 5,000 g for 5 min, followed by gentle rinsing of the pellet with 4 x 1 ml

70% acetone/H2O. The pellets were then resuspended in 240 µl HEN buffer containing

1% SDS. For labeling, the blocked samples were mixed with 0.1 vol. of biotin-HPDP

(2.5 mg/ml in DMSO) and 0.1 vol. of HEN (control) or freshly prepared sodium

ascorbate in HEN buffer. Labeling reactions were performed in the dark at room

temperature for 1 h unless otherwise indicated. Indirect sunlight was introduced by

incubating the labeling reactions near a laboratory window without direct exposure to

sunlight or change in temperature.

For direct detection of protein biotinylation, 20-40 µg of each labeling reaction

was resolved by non-reducing SDS-PAGE, followed by immunoblotting with anti-biotin

antibody (1:500). To detect an individual SNO-protein from cells or lysates, the labeling

reaction was acetone-precipitated as previously described. The washed pellet was

resuspended in 250 µl HEN/10 (HEN diluted 10 fold into H2O) containing 1% SDS,

followed by addition of 750 µl neutralization buffer (25 mM HEPES, 100 mM NaCl, 1

mM ETDA, 1% Triton-X 100, pH 7.5). This material was incubated overnight at 4 0C

with 40 µl of a streptavidin-agarose slurry. The beads were washed with 4 x 1 ml with

wash buffer (neutralization buffer + 500 mM NaCl), followed by 2 x 1 ml of

neutralization buffer. The dried beads were eluted with 50 µl HEN/10 + 1% β-

mercaptoethanol at RT for 20 min. The eluted mixture was then analyzed by SDS-PAGE,

followed by immunoblotting with anti-PTP1B (1:1000) or anti-GAPDH (1:1000)

antibodies.

Thiol Measurements and GSSG Reductase Activity Assays – Thiols were assayed

by addition of one vol. of 0.1 mg/ml DTNB in HEN. Absorbances of samples and GSH

Page 34: thesis mtf 1 - Duke University

19

standards were immediately read at 405 nm. Activity of GSSG reductase was measured

as described [65] by following GSSG-dependent consumption of NADPH at 340 nm.

SNO-Photolysis – Lysates or purified proteins were incubated in a pyrex

borosilicate NMR tube about 2 cm from a 200 W Hg-vapor lamp (Hanovia) for 3 min,

which provides approximately 4 W of radiation from ~280-500 nm. During this time, the

sample temperature was unchanged. Immediately following this incubation, the samples

were subjected to the BST as described.

2.5 Specificity of the Biotin Switch Technique for the Detection of Protein-SNOs

To evaluate the specificity of the BST, we first sought to perform the assay with

an exemplary protein on which we could introduce alternative Cys modifications,

including free sulfhydryl (-SH), S-glutathionyl mixed disulfide (-SSG),

sulfenic/sulfinic/sulfonic acids (-SOH, -SO2H, -SO3H) and S-nitrosothiol (-SNO). Human

protein tyrosine phosphatase 1B (PTP1B), with its well-characterized redox-regulated

active site Cys residue [66], satisfied these criteria. Treatment of reduced PTP1B with

H2O2 resulted in the partially DTT-reversible inhibition of enzyme activity suggestive of

a mixture of -SOH/sulfenyl amide (reversible) and -SO2H/-SO3H (irreversible)

modifications (Figure 5A). Incubation of PTP1B with oxidized glutathione (GSSG) or S-

nitrosocysteine (CysNO) resulted in S-glutathionylation (Figure 5B) or S-nitrosylation

(Figure 5C), respectively. Reduced and redox-modified PTP1Bs were then subjected to

the BST with various concentrations of ascorbate. Under these conditions, only the S-

nitrosylated PTP1B was biotinylated (Figure 5D). It should be noted that even with 100

mM ascorbate, the assay specificity was not compromised. The use of 100 mM ascorbate

Page 35: thesis mtf 1 - Duke University

20

did not, however, improve sensitivity (i.e. result in increased biotinylation of SNO-

PTP1B) relative to 10 mM ascorbate, suggesting that under these conditions the reaction

with ascorbate is not a limiting factor.

To investigate the specificity of the assay in situ, and in particular to determine

whether any cellular components might limit the specificity of the assay, we subjected

H2O2-, GSSG- and CysNO-treated lysates from HEK-293 cells to the BST and assayed

for the biotinylation of endogenous PTP1B by streptavidin pull-down and

immunoblotting (Figure 5E). As with purified PTP1B, biotinylation was observed only in

the CysNO-treated samples, but in this case the labeling of PTP1B was higher at 100 mM

than 10 mM ascorbate. This can be well rationalized: since the reaction rates for

transnitrosylation between SNOs and ascorbate vary over three orders of magnitude and

are first order for each reactant [67], low abundance protein SNOs may require larger

ascorbate concentrations to facilitate this reaction. Another possibility is that ascorbate is

consumed by oxidants present in the cellular extracts. Thus, whereas 1 mM ascorbate is

sufficient for detection of some S-nitrosylated proteins as originally described [50], the

general sensitivity of the assay may be enhanced by the use of higher ascorbate

concentrations.

To more generally assess assay specificity, we examined biotinylation by the BST

of multiple S-nitrosylated, S-oxidized, or S-glutathionylated cellular proteins. HEK-293

cell lysates or intact cells were treated with CysNO, H2O2 and either GSSG (lysates) or

diamide, a thiol-oxidizing agent (intact cells), and then analyzed by the BST. Under

these conditions, protein biotinylation was strictly dependent upon treatment with CysNO

and the inclusion of ascorbate during the labeling step (Figures 6A and 6B). To

Page 36: thesis mtf 1 - Duke University

21

complement these results, we also show that S-glutathionylated proteins were not reduced

in the BST (Figure 6C).

2.6 Conditions that Promote S-Nitrosothiol-Independent Protein Biotinylation

Although the BST appeared to be highly specific for protein SNOs under our

assay conditions, we nonetheless sought to investigate the recent claim of “artifactual”

ascorbate-dependent biotinylation of reduced BSA [59]. We reasoned that this SNO-

independent, ascorbate-dependent protein biotinylation must involve reduction of protein

disulfides (including S-methythiolated cysteines) or biotin-HPDP. Ultraviolet radiation is

known to facilitate the indirect reduction of protein disulfides [68, 69] as well as the one

e- oxidation of ascorbate to the reactive semidehydroascorbate radical, which can reduce

non-biological (reactive) aryl disulfides such as 4,4’-dithiopyridine [70]. We further

reasoned that ascorbate would rapidly reduce protein thiyl radicals [71] generated by UV

radiation, providing thiol substrate for biotinylation. Accordingly, we sought to determine

whether light could promote artifactual biotinylation during the labeling step of the BST

(the assay, as originally described, was performed in the dark).

Interestingly, pre-reduced BSA, lactoglobulin, and GAPDH were biotinylated

when the labeling reactions were exposed to indirect sunlight, but not when kept in the

dark, even after a 16 h incubation (standard incubations are 1-2 h) (Figure 7A). Similar

results were seen without DTT pre-reduction (data not shown). Further, artifactual

biotinylation was fully reversed by excess DTT (Figure 7C), confirming that the biotin

was attached to the protein via a disulfide bond. In addition, the samples exposed to

Page 37: thesis mtf 1 - Duke University

22

sunlight exhibited a marked yellowing indicative of ascorbate oxidation to

dehydroascorbate, whereas those kept in the dark remained essentially colorless.

To determine the effects of this artifact under typical laboratory conditions, we

performed the BST on reduced and S-nitrosylated GAPDH and exposed the labeling

reactions to either fluorescent lighting (in a windowless room) or indirect sunlight (a

laboratory bench approximately 2 meters from a window) (Figure 7B). Under fluorescent

lighting, only SNO-GAPDH underwent ascorbate-dependent biotinylation (Figure 7B,

lanes 1-4); this specific SNO signal was comparable to that obtained in the dark (not

shown). In contrast, indirect sunlight led to increases in ascorbate-dependent biotinylation

of both reduced and S-nitrosylated GAPDH (Figure 7B, lanes 5-8). Even in the absence

of ascorbate, indirect sunlight potentiated the biotinylation of SNO-GAPDH (Figure 7B,

lanes 3 vs. 7), consistent with known UV sensitivity of the SNO. Furthermore, the

combination of indirect sunlight and ascorbate dramatically increased the biotinylation of

SNO-GAPDH (Figure 7B, lanes 4 vs. 8), notwithstanding loss of SNO-specificity under

these conditions.

To understand the mechanism of this artifact, we examined the ability of

ascorbate to reduce biotin-HPDP to biotin-SH under the conditions of the BST.

Interestingly, nearly quantitative reduction of biotin-HPDP occured in the presence of

ascorbate and indirect sunlight, but not ascorbate alone (Figure 7D). The production of

high micromolar biotin-SH during the BST would lead to artifactual protein biotinylation

via thiol/disulfide exchange with “blocked” (i.e. S-methylthiolated) proteins.

Page 38: thesis mtf 1 - Duke University

23

2.7 UV Pre-Photolysis as a Control for the Detection of Protein S-Nitrosothiols

Under some experimental conditions, ascorbate-dependence for biotinylation may

be the primary control for confirming an endogenous protein SNO by the BST. For

example, numerous experiments have demonstrated endogenous constitutively S-

nitrosylated proteins that are largely insensitive to NOS inhibitors [25, 72-74]. To

complement the ascorbate control in instances such as these, we sought to introduce a

separate method to independently confirm the existence of an SNO moiety. Since UV

light will efficiently cleave (photolyze) an SNO [75], we reasoned that it should eliminate

SNOs. We thus exposed SNO-GAPDH to UV radiation prior to the BST. This treatment

resulted in a significant decrease in ascorbate-dependent biotinylation of SNO-GAPDH

(Figure 8A). The generality and biological utility of this approach were confirmed using

CysNO-treated HEK-293 cells (Figure 8B), which also demonstrated a strong attenuation

of biotinylation when the lysates were first subjected to photolysis.

2.8 Effects of Oxidative Stresses on Protein S-Nitrosylation

Despite reports that GSH depletion [76-78] and exogenous peroxides [79]

facilitate S-nitrosylation, relatively little is known about the overall role of oxidative

stimuli in modulating protein SNOs. Furthermore, such studies require methodologies

that can differentiate SNOs from other cotemporaneous Cys-based redox modifications.

The BST coupled with SNO-photolysis would seem to be ideal for such analyses.

To this end, we examined the effects of BCNU (carmustine)—a clinically

employed alkylating/carbamoylating agent often used to treat astrocytomas—on general

Cys-redox status and S-nitrosylation. Consistent with previous reports [80, 81], BCNU

Page 39: thesis mtf 1 - Duke University

24

efficiently inhibited GSSG reductase (GR) in HEK-293 cells (Figure 9A), which has been

shown to increase the cellular GSSG/GSH ratio and thereby induce oxidative stress [82].

Furthermore, these conditions resulted in a loss of cellular protein thiols (Figure 9B),

demonstrating the effects of BCNU on Cys-redox status of both low molecular weight

thiols and proteins. Next, HEK-293 cells stably expressing nNOS were treated with

BCNU and subjected to the BST. A dose-dependent increase in SNO-GAPDH was

observed (Figure 9C), which correlated with both GR inhibition (Figure 9A) and protein

thiol oxidation (Figure 9B). To confirm that modification of GAPDH by S-nitrosylation

was consequent upon BCNU-induced oxidative stress (and hence to differentiate SNO

from other Cys-redox modifications), we employed SNO-photolysis prior to the BST. As

expected, UV pre-photolysis of samples resulted in a marked decrease in biotinylated

GAPDH (Figure 9D).

2.9 Discussion

Jaffrey and Snyder have provided unequivocal demonstrations of the utility of the

BST for identifying SNOs [27, 50, 83-85]. Given the widespread use of the assay, it

remains important to show that assay specificity, which has to-date been examined for a

limited number of proteins [50, 85], in fact captures the general behavior of SNO-

proteins in complex biological systems. Here, we confirm that the BST is in fact highly

selective for S-nitrosylated vs. S-oxidized (sulfenic acid and disulfide) groups in proteins,

and that assay specificity, which is a unique feature of the chemistry between ascorbate

and SNO, can be generalized to many classes of proteins.

Page 40: thesis mtf 1 - Duke University

25

Interestingly, the one e- reduction potential of GSH/GSNO is ~ -180 mV [86],

indicating that ascorbate cannot directly reduce the S-NO bond (the

ascorbate/semidehydroascorbate radical Ered is +280 mV) [87]. Yet the BST requires that

ascorbate liberate NO from SNO, which constitutes a formal one e- reduction. The

solution to this paradox is to understand that ascorbate does not directly donate an

electron to SNO, but rather undergoes transnitrosation by SNO (a bimolecular reaction

involving NO+ transfer), to generate O-nitrosoascorbate, which rapidly homolyzes to

yield the semiDHA radical and NO [67]. Since this reaction with ascorbate is unique to

SNO among Cys oxidation products, it confers specificity to the BST. Our experimental

evidence strongly supports this conclusion.

Our efforts to investigate the reports of false-positive biotinylation [56-59], which

we had not previously observed even at high ascorbate concentrations [55, 64], revealed

that indirect sunlight is the likely cause. The semidehydroascorbate radical, an

intermediate in the oxidation of ascorbate to dehydroascorbate by sunlight, probably

plays a role. Though semidehydroascorbate cannot drive the single e- reduction of

biological Cys disulfides (Cys disulfide radical/disulfide Ered is ~ -1500 mV [88]), it has

been shown to reduce reactive synthetic disulfides such as 4,4’-dithiopyridine [70], which

are similar to biotin-HPDP. Indeed, we show that ascorbate in the presence of indirect

sunlight reduces biotin-HPDP to biotin-SH, which would inevitably incorporate into

protein disulfides to generate artifactual (i.e. SNO-independent) S-biotinylation,

consistent with our findings and those of Huang et al. [59]

Alternatively the data might be explained by UV light induced tryptophan

photoexcitation, enabling the transfer of one e- to an adjacent disulfide to generate a

Page 41: thesis mtf 1 - Duke University

26

disulfide radical [68, 69], which transiently dissociates into free thiol plus thiyl radical.

Thermodynamically, ascorbate should drive this reaction by further reducing Cys thiyl

radical to thiol (Cys thiol/thiyl radical Ered +920 mV [89]), which would then undergo

biotinylation. Of relevance, Walmsley et al. applied this rationale to explain their

discovery that ambient daylight reduced DTNB in the presence of ascorbate [90].

Whatever the exact sequence of electron transfer that creates the artifact, we can

recapitulate the result with indirect sunlight and ascorbate.

Interestingly, our experiments demonstrate that, even in the absence of ascorbate,

indirect sunlight promotes the biotinylation of an SNO during the labeling step of the

BST. Indirect sunlight likely contains sufficient UV light for SNO homolysis (λmax 334

nm). Though it is generally thought that thiyl radicals proportionate to disulfides or are

further oxidized, our experiments and those of others [68, 69] suggest that protein Cys

thiyl radicals are, at least in small part, reduced to thiols that can subsequently react with

biotin-HPDP. That ascorbate potentiates the effects of indirect sunlight on SNO

biotinylation would be consistent with ascorbate-mediated reduction of thiyl radical (as

described previously). Alternatively, the semidehydroascorbate radical

(semidehydroascorbate/dehydroascorbate Ered -174 mV [87]) may directly reduce an

SNO to thiol (they have similar reduction potentials; see above). Increased sensitivity of

the BST as a result of sunlight, however, comes at the expense of specificity; the labeling

step should therefore be performed under fluorescent light or in the dark.

A number of anticancer agents employed in clinical practice create an oxidative

stress. Assessment of protein oxidation vs. S-nitrosylation under these conditions has

been a challenge. The introduction of the BST coupled to SNO-photolysis has allowed us

Page 42: thesis mtf 1 - Duke University

27

to rigorously demonstrate that pharmacologically induced oxidative stress in fact

promotes GAPDH S-nitrosylation. Mechanisms for this increase in SNO-GAPDH may

include a loss of reduced cellular GSH or oxidation of proteins necessary for the

breakdown of S-nitrosothiols. Given the well-established role of SNO-GAPDH in

mediating cellular death [27, 28, 83], our data raise the idea that clinically employed

cancer chemotherapies (e.g. BCNU) may execute their functions via S-nitrosylation of

specific cytotoxic effectors (e.g. SNO-GAPDH).

Unlike previous studies that have induced oxidative stresses by inhibition of GSH

synthesis, we demonstrate that GR-dependent recycling of GSSG to GSH plays a role in

protein S-nitrosylation. GSH will accept NO from protein SNOs to generate GSNO,

which can then be metabolized by GSNO reductase [46, 47]. The mechanism by which

BCNU increases SNO-GAPDH most likely involves a relative depletion of reduced GSH

(via GR inhibition), thus abrogating the transnitrosylative pathways that generate GSNO.

In identifying GR as a regulatory component in SNO homeostasis, we establish that

oxidative stimuli may elicit their effects through S-nitrosylation, rather than exclusively

by Cys-oxidation.

Notwithstanding the reproducibility and specificity of the BST for the detection of

SNO proteins, we show here that SNO-photolysis prior to the assay may be useful as a

separate control. Whereas ascorbate generates a SNO-dependent signal, photolysis

eliminates it. Photolysis may thus find wide application in instances where the BST does

not respond to NOS agonists or inhibitors (which increase or eliminate endogenous SNO

signals) or under conditions of diverse redox stresses (e.g. BCNU). The BST (when

performed protected from sunlight) combined with SNO-photolysis should provide a

Page 43: thesis mtf 1 - Duke University

28

highly specific method for future investigations of S-nitrosylation, particularly under

conditions that require differentiation of SNOs from other Cys-based modifications.

Page 44: thesis mtf 1 - Duke University

29

Figure 5. Specificity of the BST in vitro and in lysates: analysis of protein tyrosine phosphatase-1B (PTP1B). PTP1B (4.5 mg/ml) was treated with 0.2 mM H2O2, 0.2 mM CysNO or 10 mM GSSG in HEN buffer, chemically characterized, and subjected to the BST. (A) H2O2-treated PTP1B was incubated in phosphatase assay buffer and optical absorbance at 405 nm was recorded. (B) GSSG-treated PTP1B was analyzed by anti-GSH immunoblot. (C) CysNO-treated PTP1B was subjected to the Saville assay and optical absorbance at 540 nm was recorded. (D) Each oxidatively modified or S-nitrosylated PTP1B (50 µg) was subjected to the BST with 0, 10 and 100 mM Asc. E, HEK-293 lysates (0.5 mg each) were treated with 0.5 mM H2O2, 10 mM GSSG, or 0.5 mM CysNO and subjected to the BST. Biotinylated PTP1B was detected by streptavidin pulldown and anti-PTP1B immunoblotting.

Page 45: thesis mtf 1 - Duke University

30

Figure 6. Specificity of the BST in vivo: general analysis. (A) Lysates from HEK-293 cells (0.5 mg each) were treated with 0.5 mM H2O2, 10 mM GSSG or 0.5 mM CysNO, subjected to the BST, and biotinylation of samples was assessed by anti-biotin immunoblotting. (B) HEK-293 cells were either untreated or incubated with 1 mM H2O2, CysNO or diamide for 5 min, washed and lysed in HEK-293 lysis buffer. The extracts (0.5 mg each) were subjected to the BST with or without 100 mM Asc and biotinylation of each sample was quantified by anti-biotin immunoblotting. (C) Samples from panel A were probed with anti-GSH antibody.

Page 46: thesis mtf 1 - Duke University

31

Figure 7. Indirect sunlight is responsible for ascorbate-dependent “artifactual” biotinylation. (A) BSA, lactoglobulin and GAPDH (5 mg/ml each) were pre-treated with 10 mM DTT for 30 min at RT, then desalted. The reduced samples (1 mg each) were subjected to the BST and labeling reactions were performed in the presence or absence of 20 mM ascorbate and/or indirect sunlight from a nearby window for 16 h. (B) GAPDH was biotinylated in the presence of ascorbate/sunlight and treated with 10 mM DTT prior to SDS-PAGE to identify the nature of the protein-bound biotin linkage. (C) GAPDH (5 mg/ml) was left untreated or allowed to react with 0.5 mM CysNO prior to performing the BST in the presence or absence of 20 mM ascorbate under either fluorescent light or indirect sunlight. In each case, protein biotinylation was assessed by anti-biotin immunoblotting. (D) Biotin-HPDP (0.4 mM) and ascorbate (20 mM) were incubated in HENS buffer under conditions identical to the BST labeling reaction. At indicated times, an aliquot was removed and assayed for thiol by DTNB assay.

Page 47: thesis mtf 1 - Duke University

32

Figure 8. UV pre-photolysis prior to the BST eliminates SNO-dependent signals. (A) GAPDH (5 mg/mL) was untreated or incubated with 0.5 mM CysNO. After removing excess CysNO, one sample of SNO-GAPDH was incubated for 2 min near a Hg-vapor lamp. The samples (1 mg each) were then subjected to the BST in the presence or absence of 50 mM ascorbate and biotinylation was measured by anti-biotin immunoblotting. (B) HEK-293 cells were either untreated or incubated with 0.5 mM CysNO for 5 min, followed by lysis with or without later exposure for 3 min to a UV-Vis Hg-vapor lamp. The BST was performed in the presence or absence of 50 mM ascorbate, and biotinylation was analyzed by anti-biotin immunoblotting.

Page 48: thesis mtf 1 - Duke University

33

Figure 9. S-nitrosylation of GAPDH is promoted by BCNU-induced oxidative stress. (A) WT HEK-293 cells were treated with BCNU for 6 h, lysed and assayed for GSSG reductase (GR) activity. (B) Lysates from (A) were desalted and protein thiols quantified by DTNB assay. (C) HEK-293 cells stably expressing nNOS (nNOS-HEK) were treated with BCNU for 6 h and subjected to the BST, followed by immunoblotting for GAPDH. (D) nNOS-HEK cells were either untreated or incubated for 6 h with 100 µM BCNU. Lysates were either untreated or SNO-photolyzed and subjected to the BST.

Page 49: thesis mtf 1 - Duke University

34

Chapter 3. Sulfur-Based SNO Detection: The Biotin Switch Technique

This chapter is based on a presentation I delivered to the Society for Free Radical

Biology and Medicine’s 14th Annual Meeting in Washington, DC. Following this

presentation, we were invited by Henry Forman (Associate Editor of Free Radical

Biology and Medicine) to submit a review article on protein-SNO detection. This chapter

is based largely on this review article, which aims to provide a methodological

framework for performing the biotin switch technique (BST).

Forrester, M.T., Foster, M.W., Benhar, M. & Stamler, J.S. Detection of protein S-nitrosylation with the biotin-switch technique. Free Radic Biol Med 46, 119-126 (2009).

Page 50: thesis mtf 1 - Duke University

35

3.1 Summary

Protein S-nitrosylation, the post-translational modification of cysteine thiols to

form S-nitrosothiols, is a principle mechanism of nitric oxide-based signaling. Studies

have demonstrated myriad roles for S-nitrosylation in organisms from bacteria to humans,

and recent efforts have greatly advanced our scientific understanding of how this redox-

based modification is dynamically regulated during physiological and pathophysiological

conditions. The focus of this chapter is to introduce and review the biotin switch

technique (BST), which has become a mainstay assay for detecting S-nitrosylated

proteins in complex biological systems. Potential pitfalls and modern adaptations of the

BST are discussed, as are future directions for this assay in the burgeoning field of

protein S-nitrosylation.

Page 51: thesis mtf 1 - Duke University

36

3.2 NO- vs. Sulfur-Based Assays of S-Nitrosylation

The sulfur-nitrogen bond of an SNO is particularly labile and can undergo both

homolytic and heterolytic cleavage reactions [91, 92]. The lability of the S-NO bond has

served as the cornerstone for numerous SNO detection strategies, though the chemistries

employed following SNO cleavage differ greatly between assays (Figure 10). Most

techniques detect the NO or nitrite (NO2-) liberated upon S-NO cleavage, and hence can

be considered “NO-based” strategies. In these assays, divalent mercury (e.g. HgCl2) is

often employed to heterolytically cleave the S-NO bond, producing a mercury-thiol

complex and nitrosonium ion (NO+); the latter is a potent nitrosant and undergoes rapid

hydration to NO2- at neutral pH. Techniques (spectrophotometric or fluorescent) that

detect the NO2- product include the Saville [93-95], diaminonapthalene [93, 96] and

diaminofluorescein assays [96-99].

Another common NO-based technique employs homolytic or reductive conditions

to cleave the S-NO bond, followed by chemiluminescent detection of the liberated NO

via reaction with ozone. Such methods include Hg-coupled photolysis-

chemiluminescence [75, 100] and the copper-cysteine-carbon monoxide (3C) assay [101-

103]. Though each of these NO-based methods is well suited for SNO quantitation

(relative to SNO standards), they have limited use in functional studies of S-nitrosylated

proteins within complex mixtures because the proteins of interest must be purified (e.g.

by immunoprecipitation) prior to SNO measurement. While this method has been applied

successfully in a number of cases—including S-nitrosylated hemoglobin [104-106],

caspase-3 [25, 107], thioredoxin-1 [108], c-Jun N-terminal kinase [109], G-protein-

Page 52: thesis mtf 1 - Duke University

37

coupled receptor kinase 2 [110], ryanodine receptor [111, 112] and prokaryotic OxyR

[113]—the arduous nature of the approach has limited its application.

As scientific interest in protein S-nitrosylation continues to intensify, an

increasing number of studies are relying on the biotin switch technique (BST) for the

detection of endogenously S-nitrosylated proteins (protein-SNOs). The introduction of

this assay by Jaffrey et al. in 2001 [50] has served as an impetus for studies probing S-

nitrosylation in vivo, largely due to its superb compatibility with ubiquitous molecular

methods (e.g. SDS-PAGE, immunodetection, mass spectrometry). In contrast to NO-

based assays, the BST is unique in that it targets the sulfur atom of an SNO without

regard for the fate of any liberated NO species; it can thus be considered a “sulfur-based”

strategy. As the BST employs covalent “tagging” of protein-SNOs, it can detect

individual protein-SNOs in a complex mixture (since the tag is added to the protein of

interest). For studies of a specific protein or class of proteins, the BST has therefore

proved to be of great utility (> 200 publications to date).

As described above, the major differences between NO- and sulfur-based

strategies for SNO detection are summarized in Figure 10. Importantly, the type of

information gained from either approach is different. NO-based assays are capable of

absolute quantitation (e.g. 10 nmol SNO per 1 mg protein), though they do not generally

discriminate the source of each protein-SNO. In contrast, the BST tends to be more

qualitative (relatively quantitative); its great advantage is its ability to detect individual

protein-SNOs in a complex mixture, as well as to identify novel protein-SNOs.

Page 53: thesis mtf 1 - Duke University

38

3.3 Workflow of the Biotin Switch Technique

As illustrated by Figure 11, the BST consists of three principal steps: 1)

“blocking” of free cysteine thiols by S-methylthiolation with MMTS (a reactive

thiosulfonate); 2) conversion of SNOs to thiols via transnitrosation with ascorbate; and 3)

in situ “labeling” by S-biotinylation of the nascent thiols with biotin-HPDP, a reactive

mixed disulfide of biotin. The degree of biotinylation (and thus S-nitrosylation) is

determined by either anti-biotin immunoblotting or streptavidin pull-down followed by

immunoblotting for the protein(s) of interest.

The blocking step is typically initiated by the addition of SDS and MMTS,

followed by heating at 50 °C and separation of proteins from excess MMTS by acetone

precipitation. The combination of heat and SDS functions to fully denature proteins, thus

granting MMTS optimal access to natively buried protein thiols. Effective blocking of

free thiols is required to minimize “background” biotinylation (i.e. biotinylation that

results from incomplete blocking) and to maximize assay sensitivity.

The second step in the BST converts the SNO to a free thiol so that it can be

biotinylated. This is achieved by allowing ascorbate—a di-enol antioxidant with unique

reactivity towards SNOs—to undergo a transnitrosation reaction with the protein-SNO

(Figure 12). As initially reported by Holmes and Williams [67], this reaction proceeds via

a concerted nitroso transfer mechanism whereby the SNO and ascorbate are converted to

thiol and O-nitrosoascorbate. Notably, ascorbate is several orders of magnitude more

reactive than other N- or O-based nucleophiles (e.g. amines, phenols) [114, 115]. The

reaction between SNOs and ascorbate also exhibits a critical pH-dependence and is

favored by transition metal chelation: alkaline conditions promote the nitroso transfer

Page 54: thesis mtf 1 - Duke University

39

reaction, likely by increasing the fraction of mono-deprotonated (i.e. nucleophilic)

ascorbate, whereas transition metals may enable alternative chemistry (see below). The

specificity of the BST for protein-SNOs is predicated on the fact that ascorbate will

convert SNOs to free thiols without reducing other cysteine-based oxidations such as S-

glutathionylation or S-oxides (sulfenic, sulfinic and sulfonic acids). This specificity is

supported by thermodynamic measurements (discussed below) and by several

experimental validations [50, 116] (See Chapter 3).

The third step of the BST involves biotinylation of the nascent thiol (i.e. previous

SNO site). This is performed concomitantly with the ascorbate reaction, such that the

newly liberated free thiols are immediately biotinylated. Excess biotin-HPDP is removed

by acetone precipitation, and protein biotinylation (i.e. previous S-nitrosylation) is

assessed by one of multiple routes. Avidin-agarose is frequently employed to enrich the

biotinylated proteins, followed by elution and SDS-PAGE. It is important to note that the

biotin tag is attached via a disulfide linkage, and therefore reductants (e.g. DTT) will

remove the tag. This reductive elution strategy is frequently employed to elute

biotinylated proteins from avidin-agarose prior to SDS-PAGE, especially when the

biotin-tag is not employed for immunodetection (e.g. when blotting an individual

protein).

3.4 Pre-photolysis as an Independent Strategy in the BST

There is little doubt that the signals arising from a BST following treatment with

CysNO represent bona fide S-nitrosylated proteins. Like all other biological assays,

however, internal and external controls should be employed to verify data obtained with

Page 55: thesis mtf 1 - Duke University

40

the BST, particularly when assaying endogenously S-nitrosylated proteins. These

approaches frequently include: 1) omission of ascorbate during the labeling step of the

BST; 2) overexpression or knockdown of a particular NOS isoform; and/or 3)

pharmacological activation or inhibition of NOS. Though each of these approaches has

been consistently applied to assay SNOs by the BST, we recently introduced an

independent technique that involves photolysis of the SNO prior to performing the BST

(dubbed “pre-photolysis”) [110, 116]. This methodology—a UV-based homolysis of the

S-NO bond—is a simple and specific tool for assaying SNOs with the BST (Figure 13),

particularly under conditions where NOS activation or inhibition are either unfeasible

(e.g. human patient samples) or ineffective (e.g. NOS activity-independent, a

characteristic of signaling activated by denitrosylation). This approach is based on Hg-

coupled photolysis-chemiluminescence, in which the Hg-displaceable pool of NO groups

(i.e. SNO) is photolyzed and detected via reaction with ozone. Photolysis eliminates

SNO. Consequently, pre-photolysis serves as a complementary approach to the BST.

As demonstrated in Figure 14A, the BST readily detects S-nitrosylated GAPDH in

cytokine-stimulated macrophages, in which inducible nitric oxide synthase (iNOS) is

upregulated [6, 117, 118]. Also shown in Figure 14A, the signal from endogenously S-

nitrosylated GAPDH is attenuated by a brief exposure to high intensity UV light. A

schematic illustration in Figure 13 shows how pre-photolysis and ascorbate can be

combined to rigorously assess a complex sample for protein S-nitrosylation.

Alternatively, total protein-SNOs can also be assessed by non-reducing SDS-PAGE and

immuno-detection of the biotin tag. As shown in Figure 9B, this approach easily detects

total protein-SNOs in HEK293 cells following treatment with S-nitrosocysteine (CysNO).

Page 56: thesis mtf 1 - Duke University

41

Collectively, these examples illustrate the utility of the BST for assessing S-nitrosylation

in vivo, particularly when used alongside the methodological, pharmacological and

genetic tools illustrated in Figures 13 and 14.

3.5 Design of the Pre-photolysis Apparatus

Pre-photolysis is performed by exposing the sample(s) to a strong UV radiation

source within a light-sealed reaction chamber. Overall such an apparatus should: 1)

deliver high intensity radiation at 335 nm (the ideal wavelength for S-NO photolysis [75,

119, 120]); 2) prevent heating of the sample to avoid thermolytic reactions; 3) allow

multiple samples to be photolyzed simultaneously; and 4) protect the operator from any

harmful UV exposure. We employ a 200 watt, 1.9 amp Hg vapor lamp (Ace Glass,

catalog # 7825-32), which delivers approximately 50% of its energy in the ultraviolet

region. This lamp, along with an appropriate power supply, can be fitted into a sealed

photochemical reaction cabinet (e.g. Ace Glass, catalog # 7836-20). Samples are

generally photolyzed in borosilicate glass vessels (Pierce catalog # 13504). Typically a

two-minute exposure approximately 3 cm from the lamp is sufficient to photolyze at least

50% of a protein-SNO.

Several more economical sources of UV radiation may also suffice for pre-

photolysis, though they are likely less efficacious than a Hg vapor lamp. These include a

UV transilluminator (often used to visualize DNA with ethidium bromide staining) and a

UV crosslinker that provides radiation with wavelengths greater than 300 nm. In the latter

case, a suitable apparatus would be the Ultraviolet Products (UVP) CL-1000 crosslinker

Page 57: thesis mtf 1 - Duke University

42

(UVP, catalog # 81-0112-01) fitted with five long-range UV lamps (UVP, catalog # 34-

0006-01).

3.6 Methodological Issues and Concerns Raised about the BST

Despite its widespread use, the BST is technically challenging and labor

intensive. Since each step contains potential sources error, a rigorous experimentalist will

include proper negative and positive controls to add confidence to their results. In the

case of S-nitrosylation, numerous tools can be employed to check assay efficiency and to

manipulate the stability or formation of protein-SNOs, thus adding great confidence to

results obtained with the BST.

When comparing protein-SNOs across multiple conditions, it is imperative that

each sample contains equivalent protein abundance, or “inputs”. While some degree of

normalization can be allowed, linearity (with respect to protein input and biotinylation) of

the BST has not been demonstrated. Since proteins are subjected to multiple acetone

precipitation steps and pellet washes, SDS-PAGE analysis should be performed on the

biotinylated material prior to avidin pulldown. The final step of the BST is, for all intents

and purposes, an affinity pulldown, which are almost always reported in the literature as

both an “input” and a “pulldown.” The BST should be held to the same standards.

The blocking step of the BST can also present a technical challenge, as some

protein thiols can be resistant to complete blocking, resulting in high levels of SNO-

independent biotinylation. In the absence of proper controls, incomplete blocking can be

problematic since it may lead to misinterpretation of the data (i.e. “false-positive” signal).

However, if performed correctly (e.g. use of ascorbate), ineffective blocking will lower

Page 58: thesis mtf 1 - Duke University

43

assay sensitivity (increase signal to noise) rather than specificity. Conversly, reverse

strategies such as omitting ascorbate or employing pre-photolysis can be used to

determine the degree of blocking. Both of these treatments will attenuate a true SNO

signal. Further, the blocking reaction can be lengthened to improve efficiency, though

SNO stability may be compromised at 50 °C due to thermolysis.

The specificity of the BST for SNOs (versus other cysteine-based modifications)

had been questioned in several works. Here we review the actual data in these

publications, and explain why observations are more likely artifact than source of real

concern. Landino et al. suggested that ascorbate reduces tubulin disulfides, a claim the

authors use to challenge the validity of the BST [56]. However, as discussed below,

reduction of alkyl (biological) disulfides by ascorbate is highly unfavorable

thermodynamically, and an explanation for how “disulfides” were purportedly reduced

was not provided. These authors in fact employed high concentrations of peroxynitrite to

“oxidize” tubulin, though peroxynitrite also nitrosates thiols (albeit at low yields) thereby

generating SNOs [121-123]; it is likely that S-nitrosylated tubulin was generated under

these conditions. A later report by Huang et al. argued that the BST yields “artifactual”

ascorbate-dependent biotinylation of native reduced BSA [59] (i.e. SNO-independent).

The authors did not provide a mechanistic basis for their observation, and subsequent

work has reproduced the findings of Huang, which was caused by a window light artifact

(BST should be performed in the dark) [116].

Most recently Guisitarini et al. have challenged the BST mainly on the grounds

that ascorbate reduces DTNB [124]. However, this well known reaction has no bearing

on reduction of biological (alkyl) disulfides: DTNB is a highly electrophilic, aryl

Page 59: thesis mtf 1 - Duke University

44

disulfide that readily undergoes both hydrolysis and homolysis, the latter facilitated by

ascorbate [90, 125]. DTNB is therefore not representative of biological disulfides or in

vivo Cys-based oxidation products. More perplexing, however, is these authors’

suggestion that ascorbate reduces low molecular weight biological disulfides (e.g.

glutathione, homocysteine and cysteine disulfide) and protein mixed disulfides, but not

intramolecular protein disulfides (under denaturing conditions). In fact, none of these

reactions are favored thermodynmically, and there is no kinetic or thermodynamic basis

for the difference between mixed and intramolecular protein disufides under denaturing

conditions. On closer inspection, yield of reaction with cystine and glutathione disulfide

was extremely low and appeared to saturate at a few percent over several hours,

consistent with the presence of a contaminant (e.g. contaminant metals) or other artifact.

Further, the authors measured disulfide reduction in the continuous presence of

fluorescent thiol alkylators (bromobimanes), which irreversibly pull reactions by Le

Chatelier’s principle, and treated other samples with DTT to prevent “possible re-

oxidation of thiols after their reduction by ascorbate” [124]. These approaches, which

would artifactually generate thiols, cast doubt on the conclusions. These authors also do

not provide a mechanistic explanation for their results or reference prior work to the

contrary [61, 62, 89, 126, 127]. But most important of all, they do not perform the BST,

whereas a comparative analysis of the BST across cell lysates derivatized by S-

glutathionylation, S-oxidation and S-nitrosylation revealed that only the S-nitroylated

proteins were detected [116].

Shared between all these claims of “artifactual signals” is the notion that

ascorbate can reduce alkyl disulfides. However, this concept is both unsupported and

Page 60: thesis mtf 1 - Duke University

45

difficult to reconcile experimentally with the use of MMTS as a blocking agent, which

converts protein thiols to mixed methyl disulfides. Clearly the BST would never work if

ascorbate removed these “blocked” methyl disulfides. Further, the two electron standard

reduction potential of cysteine disulfides is -170 to -320 mV [89], indicating that this

endergonic reaction will not likely couple to the two electron oxidation of ascorbate,

which is also unfavorable (the standard reduction potential of dehydroascorbate to

ascorbate is +70 mV [60]). Thus, one would not expect ascorbate to directly reduce any

biological protein cysteine oxidation products (except highly transient and unstable thiyl

or sulfinyl radicals). In contrast, these electrochemical measurements favor the reverse

reaction—thiol-dependent reduction of dehydroascorbate to ascorbate—a scenario

supported by extensive in vitro and in vivo experimentation [61, 62, 127]. The same

thermodynamic arguments hold for higher S-oxides as well [89].

Although the BST is highly specific for protein SNOs [116], we have noted that

the presence of indirect sunlight (from an adjacent window) during the labeling step

closely recapitulates the artifactual results of Huang et al. [59]. This SNO-independent

biotinylation reaction is fully reversed by DTT, confirming that indirect sunlight indeed

promotes an artifactual ascorbate-dependent protein-biotin disulfide during the labeling

step. Under these conditions, biotin-HPDP (a synthetic aryl disulfide) undergoes

quantitative reduction to biotin-thiol, though ascorbate exhibited no measureable

reactivity in the absence of sunlight [116]. This production of high micromolar biotin-

thiol during the BST would lead to artifactual protein biotinylation via thiol/disulfide

exchange with “blocked” (i.e. S-methylthiolated) proteins. The mechanism of this side

reaction may involve homolysis of biotin-HPDP to generate an unstable alkyl thiyl

Page 61: thesis mtf 1 - Duke University

46

radical of the biotin group and a resonance-delocalized aryl radical of 2-thiopyridine. In

the presence of ascorbate, these thiyl radicals would be reduced to thiol [89] and

therefore contribute to the observed artifact. Importantly, similar reactions between light,

aryl disulfides and ascorbate have been reported [70, 90]. These issues underscore the

critical need to perform the labeling step of the BST in a light-free environment, and to

employ metal chelators (more below).

3.7 Potential Problems Associated with Exogenous Metals in the BST

Ascorbate will drive one-electron reductions of Cu2+ and Fe3+ ions, which may

result in reactions that compromise BST specificity, including production of ascorbate

and hydroxyl radicals. These reactions are readily mitigated through the use of metal

chelators (e.g. EDTA, DTPA, neocuproine) [128-131]. Despite the widespread

application of metal chelation with the BST, one recent report suggests a metal-

dependence for the BST to efficiently detect SNOs [132]. However, this conclusion

overlooks the different chemistry operative in the presence and absence of transition

metals: transition metals favor reductive chemistry by ascorbate over transnitrosation.

Thus at low concentrations of ascorbate (<1 mM) as employed by Wang et al. [132],

transition metals promote direct SNO reduction (metals serve as catalysts). Conversely,

transnitrosation reactions, which form the basis of BST specificity are favored by metal

sequestration and higher ascorbate concentrations [67]. Increasing the amount of

ascorbate (5 – 50 mM ascorbate) should remedy the reported “requirement” for

exogenous metals that may increase assay sensitivity, but at the expense of specificity. In

particular, redox active metals are known to promote manifold reactions (both ascorbate

Page 62: thesis mtf 1 - Duke University

47

dependent and independent), such as Cu2+-mediated thiol oxidation [133-135] and Fenton

chemistry [136-138]. Given the potential for such side reactions to compromise assay

specificity, the addition of exogenous redox active metals in the BST cannot be

advocated without more rigorous validation in complex biological systems.

Given the known reactivity of Hg towards thiols and SNOs, numerous studies

have attempted to employ Hg salts (e.g. HgCl2) to quench SNOs prior to performing the

BST (and thus serve as an internal assay control in lieu of omitting ascorbate). Despite

the frequent application of Hg salts in NO-based assays (described above), the high

affinity of Hg salts for free thiols [139-141] makes this strategy inherently problematic

for the BST. As nicely demonstrated by Zhang et al. [142], HgCl2 prevents biotinylation

of free thiols and lowers any signal irrespective of S-nitrosylation status. Since all sources

of biotinylation (i.e. signals) in the BST are attenuated by Hg salts, it is unclear how such

a strategy could be used to distinguish an SNO from a false positive signal due to

incomplete blocking; therefore, Hg salts should not be used in lieu of other well-validated

controls (e.g. pre-photolysis, NOS inhibition, omitting ascorbate). By contrast, in an NO-

based assay such as photolysis-chemiluminesence, Hg-thiol reactions identify the source

of NO with Cys thiol by quenching the signal.

3.8 Modern Permutations and Adaptations of the BST

Various groups have adapted the BST to suit specific experimental goals. For

example, several labs have trypsinized the biotinylated material prior to avidin pulldown

to allow mass spectrometry-based determination of protein-SNO sites (i.e. specific

cysteine residues targeted by S-nitrosylation) [110, 143-145]. In addition, a thiol-reactive

Page 63: thesis mtf 1 - Duke University

48

hexahistidine tag has been used in lieu of biotin-HPDP to further facilitate mass

spectrometric analysis [146]. These studies have further demonstrated the utility of the

BST for SNO-oriented proteomics and bioinformatics. However, some of these strategies

rely on a single peptide (containing the SNO site) for the identification of an entire

protein, and thus stringent mass spectrometric criteria must be applied to these

approaches to avoid incorrect protein (and SNO site) identification.

In another novel approach, Kettenhofen et al. substituted biotin-HPDP with thiol-

reactive cyanine dyes to combine in-gel fluorescent detection strategies with the BST

[147, 148]. Han et al. similarly employed coumarin-based fluorophores to the same end

[149]. A major advantage of this approach is that it is readily adaptable to 2-dimensional

differential in-gel electrophoresis (2D-DIGE) techniques where multiple differentially

fluorescent-labeled samples are co-analyzed on a single 2D electrophoresis gel [150,

151].

3.9 Sample Protocol for the Biotin Switch Technique

Overall steps of the BST:

1) Blocking

2) Labeling

3) Pulldown

4) SDS-PAGE and/or immunodetection

Page 64: thesis mtf 1 - Duke University

49

General considerations for the BST:

A) A wide range of protein amount can be utilized (generally 0.3 to 5 mg of total protein

per sample).

B) A sensitive antibody with low background signal is critical for detecting an individual

protein-SNO within a lysate.

C) As discussed above, positive and negative controls should be included in every BST.

For example, addition of 0.1 mM CysNO to one sample (prior to the BST) is an optimal

positive control. A non-ascorbate or photolyzed sample serves as a convenient negative

control.

D) Acetone precipitations are typically used to remove excess reagents and isolate

proteins. However, for small sample volumes (< 100 µl) or protein quantities (< 300 µg),

acetone precipitation is often inefficient or inconsistent. Under these conditions, P-6

desalting gel (Bio-Rad) or G-25 sephadex (Amersham) can be substituted for acetone

precipitations.

Necessary reagents and notes:

Methyl methanethiosulfonate (MMTS) – Fluka, catalog # 64306.

Biotin-HPDP – Pierce, catalog # 21341.

Streptavidin agarose – Fluka, catalog # 85881.

Sodium ascorbate – Fluka, catalog # 11140. This ascorbate has been assayed for metal

content, which is reported on the bottle. We have noted that this ascorbate is considerably

more stable in aqueous solution than other preparations of ascorbate (i.e. yellowing due

to slow oxidation to dehydroascorbate is undetectable).

Page 65: thesis mtf 1 - Duke University

50

Stock solutions:

HEN buffer (100 mM HEPES, 1 mM EDTA, 0.1 mM neocuproine pH 8.0).

Alternatively, DTPA can be substituted for EDTA.

HENS buffer – HEN buffer with 1% SDS (w/v).

HEN/10 buffer – HEN buffer diluted 10-fold with dH2O.

HENS/10 buffer – HEN/10 buffer with 1% SDS (w/v)

Neutralization buffer (25 mM HEPES, 100 mM NaCl, 1mM EDTA, 0.5% Triton X-100,

pH 7.5).

Wash buffer – neutralization buffer containing 600 mM NaCl.

25% SDS in dH2O (w/v).

70% acetone in dH2O (v/v).

Solutions to prepare immediately before assay:

10% MMTS in N,N-dimethylformamide (DMF) (v/v).

200 mM sodium ascorbate in HEN buffer (stored in the dark on ice).

2.5 mg/ml Biotin-HPDP in DMSO.

Elution buffer – HEN/10 containing 1% beta-mercaptoethanol (v/v).

Sample protocol for the BST:

1) SAMPLE PREPARATION: Cells are lysed as desired. For example, a 6 cm plate

of HEK293 cells can be lysed in 0.40 ml of 25 mM HEPES, 50 mM NaCl, 0.1

mM EDTA (DTPA may be a better metal chelator; note also that neocuproine is

present in HEN buffer), 1% NP-40, 0.5 mM PMSF + protease inhibitors, pH 7.4.

Page 66: thesis mtf 1 - Duke University

51

Repeated passage through a 28-gauge needle will increase lysis efficiency. It is

critical to avoid DTT (or other reductants) in the lysis buffer because they will

destabilize SNOs and interfere with the BST. Following centrifugation, protein

concentrations are measured (e.g. BCA assay, Pierce). Samples are adjusted with

lysis buffer to achieve equal concentrations of total protein.

2) BLOCKING: For experiments employing 0.5 – 2 mg of protein, samples are

diluted to 1.8 ml with HEN buffer. Next, 0.2 ml of 25% SDS is added along with

20 µl of 10% MMTS (final 2.0 ml volume with 2.5% SDS and 0.1% MMTS).

Samples are incubated at 50 °C in the dark for 15 – 20 min with frequent

vortexing. This step is frequently performed in 15 ml conical tubes. Alternatively,

MMTS may be added directly to the lysis buffer to initiate blocking immediately

following lysis, though it should be noted that MMTS interferes with the BCA

protein assay. For some proteins, the blocking step can be efficiently performed at

room temperature (e.g. S-nitrosylated caspase-3), though this should be

determined empirically for each protein.

3) PRECIPITATION: Three volumes of cold acetone (6 ml) are added to each

sample. Proteins are precipitated for 20 min at -20 °C, and collected by

centrifugation at 2,000 g for 5 min. The clear supernatant is aspirated and the

protein pellet is gently washed with 70% acetone (4 x 5 ml).

4) LABELING: Following resuspension in 0.24 ml HENS buffer, the material is

transferred to a fresh 1.7 ml microfuge tube containing 30 µl biotin-HPDP (2.5

mg/ml). The labeling reaction is initiated by adding 30 µl of 200 mM sodium

Page 67: thesis mtf 1 - Duke University

52

ascorbate (we tend to use a final concentration of 20 mM ascorbate, though less

amounts may suffice [50] and higher amounts can be used as necessary [116]).

Alternatively, an equivalent concentration of NaCl can be used as an ascorbate-

free control. Samples are rotated at room temperature in the dark for 1 h. It is

critical to avoid any sources of sunlight during this step.

5) PRECIPITATION: Three volumes of cold acetone (0.9 ml) are added to each

sample. Proteins are precipitated for 20 min at -20 °C, and collected by

centrifugation at 5,000 g for 5 min. The clear supernatant is aspirated and the

protein pellet is gently washed with 70% acetone (4 x 1 ml).

6) PULLDOWN: Following complete resuspension in 0.25 ml HENS/10 buffer, 0.75

ml of neutralization buffer is added. A small fraction of each sample (e.g. 10 µl)

is removed for analysis of protein “input”. The remaining material is transferred

to a fresh 1.7 ml microfuge tube containing 25 – 50 µl of pre-washed avidin-

affinity resin. It is critical that bead volumes in each sample are the same. The

samples are gently rotated for 12 – 18 h at 4 °C.

7) WASHING: Avidin beads are collected by centrifugation at 200 g x 10 s in a

swinging-bucket rotor, followed by washing with wash buffer (4 x 1 ml). After

the final wash, the beads are fully dried via gentle aspiration with a 28-gauge

needle.

8) ELUTION: To each sample is added 30 – 50 µl of elution buffer. Proteins are

eluted at room temperature with frequent agitation, followed by centrifugation at

5,000 g x 30 s. Supernatant is collected without disturbing the pelleted resin, and

Page 68: thesis mtf 1 - Duke University

53

mixed with 6x Laemmli loading buffer. For analysis of protein-SNOs via

immunodetection of biotinylation, elution is performed by heating the beads to 95

°C in 30 – 50 µl of HENS/10 buffer containing non-reducing Laemmli loading

buffer (all reductants must be avoided).

Page 69: thesis mtf 1 - Duke University

54

Figure 10. A general comparison of NO- and sulfur-based strategies for detecting protein S-nitrosylation. As an example, three lysates containing various amounts of protein S-nitrosylation are subjected to both NO- and sulfur-based assays. NO-based strategies include the Saville and diaminofluorescein (DAF) assays, which employ a chemical probe, and Hg-coupled photolysis-chemiluminescence (PCL), which detects NO gas liberated by SNO homolysis and can differentiate SNO from metal-NO. Importantly, this assay is highly sensitive (low nanomolar SNO concentrations can be detected) and has been well-validated with genetic models of disrupted NO/SNO metabolism. It therefore serves as a standard method for probing S-nitrosylation in vivo. With a complex biological sample (e.g. a lysate), these NO-based strategies can readily determine the absolute amount of SNO per sample, but cannot readily detect an individual protein-SNO. A sulfur-based strategy, such as the biotin switch technique (BST), employs covalent “tagging” at the sulfur atom of each SNO, thus facilitating relative quantitation and protein-SNO identification.

Page 70: thesis mtf 1 - Duke University

55

Figure 11. Overview of the biotin switch technique (BST). In the example shown, three lysates with various degrees of S-nitrosylation are subjected to the assay. The “blocking” step involves S-methylthiolation of each cysteine thiol with S-methyl methanethiosulfonate (MMTS). Next, ascorbate is employed to convert each SNO to a free thiol via a transnitrosation reaction to generate O-nitrosoascorbate. In the “labeling” step, each nascent free thiol (previously an SNO site) is biotinylated with biotin-HPDP. Biotinylated proteins are enriched by avidin affinity media, and analyzed by SDS-PAGE/immunoblotting. Total protein-SNOs or an individual protein-SNO can be detected with avidin-HRP or with an antibody against a protein of interest, respectively. As illustrated, the degree of pulldown correlates with protein S-nitrosylation. Prior to avidin pulldown, a small fraction of each sample is analyzed to determine protein “input.”

Page 71: thesis mtf 1 - Duke University

56

Figure 12. Transnitrosation between ascorbate and an S-nitrosothiol. This reaction forms the basis of the biotin switch technique in which an SNO is selectively converted to a disulfide linked biotin moiety. In this reaction, the nucleophilic oxygen of monodeprotonated ascorbate attacks the electrophilic nitrogen of the RSNO, leading to nitroso transfer. The products are thiol and O-nitrosoascorbate, which undergoes rapid homolysis. The nitroso group is colored purple.

Page 72: thesis mtf 1 - Duke University

57

Figure 13. An illustration employing UV pre-photolysis and ascorbate controls within the BST for the analysis of a single lysate. One sample is split to two fractions, which are either untreated or exposed to a strong UV light source (“pre-photolysis”) prior to performing the BST. The pre-photolysis step leads to homolytic cleavage of the S-NO into NO and an unstable thiyl radical, which is either reduced to a thiol or oxidized to a higher S-oxide. In either case, the signal from an SNO will be attenuated by pre-photolysis, while free thiols or disulfides are unaffected. Inclusion or exclusion of ascorbate during the labeling step of the BST can also be employed to assay for protein-SNOs. This approach, which employs internal controls (e.g. pre-photolysis and ascorbate), is ideal when external controls such as NOS inhibition or activation are not feasible (e.g. human tissue samples) or ineffective (as in the case of reactions regulated by denitrosylation).

Page 73: thesis mtf 1 - Duke University

58

Figure 14. A typical BST detects both endogenous and exogenous S-nitrosylation in cultured mammalian cells. (A) Murine RAW264.7 macrophages were either untreated or cytokine-stimulated with lipopolysaccharide (500 ng/ml) and IFN-γ (100 U/ml) for 16 h, which drives NO production. Cellular extracts were subjected to the BST and probed for S-nitrosylated GAPDH (GAPDH-SNO), along with ascorbate and pre-photolysis controls. Notably, omission of ascorbate leads to nearly complete loss of biotinylation and pre-photolysis with a Hg vapor lamp greatly attenuates the same signal. (B) Whole cellular protein-SNOs are detected by treating HEK293 cells with 200 µM S-nitrosocysteine (CysNO) for 10 min. Cellular extracts were subjected to the BST and 5% of each biotinylation reaction (~ 40 µg) was analyzed by immunoblotting with avidin-HRP and anti-GAPDH antibody (for “input”).

Page 74: thesis mtf 1 - Duke University

59

Chapter 4. Proteomic Analysis of S-Nitrosylation by Resin-Assisted Capture

This chapter is based on a manuscript that is currently under review at Nature

Biotechnology.

Forrester, M.T., Thompson, J.W., Foster, M.W., Moseley, M.A., & Stamler, J.S. Proteomic analysis of S-nitrosylation by resin-assisted capture. Nat Biotech, Under Review (2009).

Page 75: thesis mtf 1 - Duke University

60

4.1 Summary

The biotin switch technique has been widely adopted for assaying protein S-

nitrosylation, the ubiquitous nitric oxide-based Cys modification. We have developed a

simple solid-phase assay that detects S-nitrosothiols by resin-assisted capture (SNO-

RAC) via a thiol-reactive sepharose matrix in lieu of thiol biotinylation and streptavidin

pulldown. Compared to the biotin switch technique, SNO-RAC requires fewer steps, is

more economical, detects high molecular weight S-nitrosylated proteins more efficiently,

and is more easily combined with mass spectrometric methods for the identification of S-

nitrosylation sites (SNO-sites) in proteins. Accordingly, we utilized SNO-RAC to

identify 91 unique SNO-sites in murine macrophages and E. coli subjected to an

exogenous nitrosative stress, as well as selected S-nitrosylated proteins that form

following cytokine stimulation of macrophages. Collectively, our results demonstrate the

utility of SNO-RAC for studying S-nitrosylation in the context of cellular biology and

proteomics. The described methodology is furthermore adapted to the study of alternative

thiol-based modifications, including protein S-acylation/palmitoylation.

Page 76: thesis mtf 1 - Duke University

61

4.2 Introduction

Nitric oxide (NO) exerts a ubiquitous influence on cellular signaling, largely via

the coordinated S-nitrosylation/denitrosylation of critical cysteine residues in proteins. It

is also increasingly evident that dysregulated S-nitrosylation plays a causal role in many

pathophysiological conditions from neurodegeneration [51, 52] to heart failure [110].

Progress in elucidating the role of S-nitrosylation in health and disease requires methods

for identification of protein-SNOs, as well as the specific Cys residues that are targets of

S-nitrosylation. To date, the biotin switch technique (BST) best satisfies these criteria. As

originally described by Jaffrey and Snyder [50], the BST consists of three principal steps:

“blocking” of free Cys thiols by S-methylthiolation with a reactive thiosulfonate,

ascorbate-dependent conversion of SNOs to thiols, and “labeling” by S-biotinylation of

the nascent thiols with biotin-HPDP, a reactive mixed disulfide of 2-thiopyridine. The

degree of biotinylation (and thus S-nitrosylation) is determined by either anti-biotin

immunoblotting or streptavidin pull-down followed by immunoblotting for the protein(s)

of interest. Efficacious blocking of free thiols is requisite for minimizing background

biotinylation (i.e. maximizing signal-to-noise), but the specificity of the BST rests upon a

unique transnitrosation reaction between ascorbate and SNOs to generate free thiol [67],

whereas reduction of other Cys-based oxidative modifications such as S-glutathionylation

or S-oxidations does not occur [116].

Although the BST has been employed to characterize SNO proteomes [50, 58,

143, 145, 152], it is labor intensive and relatively low-throughput, and thus not ideally

suited for proteomic analysis of protein-SNOs. To address these issues, we devised a

simple and robust alternative that involves resin-assisted capture of protein-SNOs on a

Page 77: thesis mtf 1 - Duke University

62

solid-phase resin, dubbed “SNO-RAC” (Figure 15). SNO-RAC consists of three

principal steps: “blocking” of free Cys thiols by S-methylthiolation with MMTS,

ascorbate-dependent conversion of SNOs to thiols and covalent “capture” of the nascent

thiols with a reactive resin-bound disulfide. Once the protein-SNOs are immobilized,

buffers are readily exchanged by washing (thus obviating any complicated dialyses or

desalting matrices).

4.3 Materials and Methods

Materials and cell culture – All reagents were from Sigma unless otherwise

specified. Monoclonal antibodies to human PTP1B, caspase-3, thioredoxin-1 and iNOS

were from BD Biosciences. Monoclonal and polyclonal antibodies to GAPDH and

peroxiredoxin-1, respectively, were from Millipore. Human embryonic kidney (HEK293)

and murine macrophage (RAW264.7) cells were cultured at 37 °C in an atmosphere of

5% CO2 and 95% air in Dulbecco’s Modified Eagle’s Medium containing 10% fetal

bovine serum, 100 U/ml penicillin, 200 µg/ml streptomycin and 1.0 mM L-Arg. Cells

were lysed in HEN buffer containing 0.2% NP-40 and 0.2 mM PMSF, centrifuged at

10,000 g x 10 min and protein was measured by the BCA assay (Pierce). Samples were

then subjected to SNO-RAC or BST as described below.

Preparation of SNO-RAC Resins – To a slurry of sepharose-NHS (Amersham

Biosciences) was added 5 volumes of 50 mM cystamine in 100 mM phosphate pH 8.0.

Following rotation for 12 h at 25 ºC, the resin was thoroughly washed with 100 mM

phosphate pH 8.0, H2O, and MeOH. The resin was reduced with 10 volumes of 100 mM

DTT in 50 mM phosphate pH 8.0, then repeatedly washed with 100 mM phosphate pH

Page 78: thesis mtf 1 - Duke University

63

8.0, followed by MeOH. Between 4 and 8 volumes of 100 mM 2- or 4-pyridyl disulfide

(PDS) in MeOH was then added to the resin, followed by rotation at 4 ºC for 12 h in the

dark. The resin was washed repeatedly with MeOH, H2O and finally isopropanol, and

stored at 4 °C in the dark. Total binding capacity was determined by treating a small

portion of the resin with DTT, followed by measuring the absorbance of 2-thiopyridone

at 343 nm (ε = 8.08 mM-1 cm-1) or 4-thiopyridone at 324 nm (ε = 19.8 mM-1 cm-1). Resin

capacity was generally 4-6 µmol disulfide/ml resin. To verify that no free thiol groups

remained (and thus that disulfide formation with 2- or 4-PDS was complete), a small

aliquot of resin was incubated with 0.1 mg/ml DTNB and absorbance was measured at

412 nm with GSH standards (no free thiol should be present). Alternatively, thiopropyl

sepharose (GE/Amersham), which is supplied with 2-PDS groups, can be used. This resin

can be functionalized with a 4-PDS group by treatment with DTT and 4-PDS, as

described above.

Preparation of S-nitrosylated GAPDH – CysNO was prepared by mixing equal

volumes of 200 mM cysteine in 1 M HCl and 200 mM NaNO2 and used immediately.

Purified GAPDH from rabbit muscle (1 mg/ml, ~ 25 µM, 0.1 ml volume) was incubated

with 0.5 mM CysNO in HEN buffer (100 mM HEPES, 1 mM EDTA, 0.1 mM

neocuproine, pH 8.0) at room temperature in the dark for 30 min. Excess CysNO was

removed by passage through P-6 desalting gel (Bio-Rad) pre-equilibrated in HEN buffer.

Detection of S-Nitrosylated Proteins by the BST or SNO-RAC – The BST was

performed as described [64, 116]. For SNO-RAC, “blocking” was as for BST, except the

resuspended pellet of “blocked” material in HENS buffer (HEN + 1% SDS) was added to

Page 79: thesis mtf 1 - Duke University

64

2- or 4-PDS sepharose (typically 200 µl HENS and 50 µl resin slurry per mg protein) in

the presence or absence of sodium ascorbate (Fluka, final 50 mM unless indicated

otherwise). Following rotation in the dark for 2-4 h, the resin was washed with 4 x 1 ml

HENS buffer, then 2 x 1 ml HENS/10 buffer (HENS diluted 1:10). For Cy3 or Cy5

labeling, samples were equilibrated in 100 mM sodium borate pH 8.5, followed by

addition of 5 µl Cy3- or Cy5-NHS (GE Life Sciences, 1 mM in DMSO), and rotated at

RT in the dark for 1 h. Captured proteins were eluted with 60 µl HEN/10 containing 100

mM 2-mercaptoethanol for 20 min at RT, and 40 µl each eluant was used for SDS-

PAGE.

Peptide-SNO Isolation by SNO-RAC – Resins containing immobilized protein(s)

were rotated at 37 ºC in the presence of 0.5 µg trypsin gold (Promega) in 1 ml of 50 mM

NH4HCO3 1 mM EDTA at 37 ºC for 12 h. When indicated, samples were then acetylated

with 5 µl H6- or D6-acetic anhydride (Sigma) in 1 ml of 100 mM sodium phosphate pH

8.0, and rotated for 30 min. Following extensive washing with 5 x 1 ml HENS/10 and 5 x

1 ml 10 mM NH4HCO3, samples were eluted for 30 min in 200 µl 10 mM NH4HCO3

with 10 mM DTT. The eluant was removed, beads were rinsed with 400 µl dH2O, and

these two fractions were combined, passed through a 0.45 µm filter and concentrated by

speedvac.

MALDI-TOF Mass Spectrometry – For MALDI, concentrated peptides were

mixed 1:1 with matrix solution (20 mg/ml α-hydroxycinnamic acid in 1:1

acetonitrile/dH2O with 0.1% TFA) and analyzed on either an Applied Biosystems

Voyager DE Pro or a Bruker Ultraflex I MALDI-TOF operating in positive-ion reflector

Page 80: thesis mtf 1 - Duke University

65

mode. Spectra were scanned from 1000 to 6000 amu and a minimum of 200 spectra were

accumulated for each sample.

Identification of high mass protein-SNOs by in-gel digestion and LC-MS/MS – All

gel manipulations were performed in a laminar flow hood. Protein bands were cut into

small pieces, washed with 500 µl 1:1 acetonitrile/100 mM ammonium biocarbonate

(AmBic) pH 8, for 15-30 minutes on a rotator or until all coomassie dye was removed

and dried at 50 ˚C. Gel pieces were swelled in 50 mM AmBic containing 10 ng/µl trypsin

(Promega, sequence grade) and for 16-18 h at 37 ˚C. Peptides were extracted with 20-30

µl 1% formic acid, 2% acetonitrile in water and vortexed for 30 min. Supernatant was

removed, concentrated by SpeedVac, resuspended in LC buffer (0.2% formic acid, 2%

acetonitrile in water) and analyzed by LC-MS/MS, using a Q-TOF Premier mass

spectrometer (Waters). Data processing and database searching were performed using

Mascot Distiller and Mascot with 20 ppm precursor and 0.04 Da product ion tolerances.

SNO-RAC of E. coli and identification of SNO-sites by isotope-encoded LC-

MS/MS – K-12 E. coli (strain MG1655) were grown to OD 1.0 in M9 minimal media

containing 0.2% glucose and left untreated or incubated with 500 µM CysNO for 5 min

at 37 °C. Extracts (10 mg protein each) were subjected to SNO-RAC as described,

trypsinized, labeled with H6- or D6-acetic anhydride and eluates were concentrated by

speedvac. Peptides were identified by LC-MS/MS on a Q-ToF Premier mass

spectrometer (Waters) coupled to a nanoAcquity liquid chromatograph (Waters). For the

LC separation, a 75 um x 150 mm Acquity BEH C18 column was used with a gradient of

3 to 40% mobile phase B over 60 min, with a flow rate of 0.4 µl/minute. Mobile phases

Page 81: thesis mtf 1 - Duke University

66

A and B used for the separation were 0.1% formic acid in water and acetonitrile,

respectively. Data-dependent acquisition (DDA) was used with 1 second survey and

MS/MS scans, and the mass spectrometer was allowed to sample the top three intensity

ions of each survey scan for MS/MS. A 200 second exclusion window was used for each

parent ion after MS/MS data collection. A lockmass calibration scan of Glu-

Fibrinopeptide B (m/z 785.8426) was introduced every 30 sec.

The DDA raw data were processed using Mascot Distiller (Matrix Sciences, Ltd)

and searched with Mascot v2.2 (Matrix Sciences, Ltd.) against the Swissprot 51.6

database, with E. coli taxonomy (www.expasy.org). The search parameters were: 1

missed cleavage, H3/D3-acetyl (on Lys residues) and H3/D3-acetyl (on N-termini) as

variable modifications, trypsin enzyme specificity, and 20 ppm precursor and 0.04 Da

product ion mass accuracy (p values < 0.01, 0.05 and 0.10 corresponded to Mascot scores

> 30, 23 and 20, respectively).

SNO-RAC of RAW264.7 macrophages and identification of SNO-sites by label-

free LC-MS/MS – Murine macrophages (RAW264.7 cells) were grown in 4 x 15 cm

dishes to ~ 80% confluence and were incubated ± 200 uM CysNO for 10 min or exposed

to 500 ng/ml LPS and 100 U/ml IFN-γ for 18 h. Cells were washed and collected with

cold PBS, lysed in 1 ml of HEN + 0.5% NP-40 and 6 mg each lysate was subjected to

SNO-RAC as described. Following on-resin trypsinization, peptides were eluted with 5

mM DTT and analyzed using a Waters MSE on a QToF Premier Mass spectrometer. Test

and control samples were analyzed three times by this method. Data was processed using

ProteinLynx Global Server (PLGS) v2.3 with the IdentityE search algorithm, using 15

ppm precursor and 30 ppm product ion tolerances. Peptides were considered as a

Page 82: thesis mtf 1 - Duke University

67

confident identification if there were 8 or more product ions matched, and these are listed

in Table 2. Also included in the table is the number of times the particular peptide was

identified out of three experiments, the maximum number of product ions observed, the

maximum mass error for the precursor ion, and the fold-change observed between test

and control (maximum value 1000).

4.4 An In Vitro Analysis of SNO-RAC with Purified GAPDH Protein

Using commercially available Sepharose-NHS 4B (Amersham/GE), we first

synthesized 2- and 4-pyridinyl disulfides of resin-bound cysteamine to test the SNO-RAC

approach. As an initial test, purified GAPDH was treated ± CysNO and subjected to

SNO-RAC. Pulldown of GAPDH was both CysNO- and ascorbate-dependent (Figure

16A), and as previously observed with the BST [116], the greater recovery of GAPDH

with 50 vs. 5 mM ascorbate suggests that reduction of SNO to free thiol is rate-limiting

under these conditions. Immobilized GAPDH-SNO was directly trypsinized “on-resin”

(Figure 16B), bound peptides were eluted with DTT and directly analyzed by MALDI-

TOF MS (Figure 16C). Tryptic peptides containing each of the protein’s four Cys

residues were readily detected, demonstrating that only Cys-containing peptides are

isolated by SNO-RAC and that each Cys residue is S-nitrosylated under these in vitro

conditions.

4.5 Assessment of Intracellular S-Nitrosylation by SNO-RAC

To examine the suitability of SNO-RAC for studying intracellular S-nitrosylation,

a side-by-side comparison of the BST and SNO-RAC using extracts from control or

Page 83: thesis mtf 1 - Duke University

68

CysNO-treated human embryonic kidney (HEK293) cells was performed. Eluates were

analyzed for total proteins-SNOs by silver staining (Figure 17A) or specific protein-

SNOs by immunoblotting for either protein tyrosine phosphatase-1B (PTP1B), caspase-3

or thioredoxin-1 (Figure 17B). Similar results were obtained with commercially available

thiopropyl-sepharose (Amersham/GE) functionalized with either 2- or 4-PDS groups

(data not shown). The sensitivity of SNO-RAC appeared superior to the BST for proteins

above ~100 kDa, perhaps due to the fewer number of required acetone precipitation

steps. To demonstrate this phenomenon at the level of proteomics, SNO-RAC-isolated

protein-SNOs with molecular mass >150 kD were excised and identified by LC-MS/MS

(Table 1). Importantly, SNO-RAC does not suffer from interference by streptavidin,

which is often co-eluted during the BST (Figure 17A, indicated by an asterisk). SNO-

RAC also avoids interference from endogenously biotinylated proteins, which is a

potential source of artifact with the BST [64] (though reductive elution should

theoretically not release endogenously biotinylated (i.e. amide-linked) peptides from

avidin-agarose). Although the BST and SNO-RAC exhibit similar sensitivities for lower

mass proteins in the general case, SNO-RAC may be the more sensitive method in

selected instances.

4.6 Assessment of Endogenous S-Nitrosylation by SNO-RAC

To gauge the suitability of SNO-RAC for the detection of endogenous protein-

SNOs, we assessed the S-nitrosylation of GAPDH, a well-established SNO-based cell

death effector [27, 83], in RAW264.7 macrophages with or without stimulation of NO

production. Although it has recently been shown that denitrosylase activities are

Page 84: thesis mtf 1 - Duke University

69

operative following cytokine stimulation (and thus whole cell SNO levels may not

directly reflect NOS activity) [47, 153], the steady state levels of GAPDH-SNO are

indeed determined by iNOS activity [27]. With SNO-RAC, GAPDH-SNO was readily

detected in cell lysates following an 18 h induction of inducible nitric oxide synthase

(iNOS) by lipolysaccharide/interferon-γ (LPS/IFN-γ) (Figure 18A). As an independent

control of assay specificity, SNO-RAC was performed on lysates after exposure to UV

radiation (“pre-photolysis”), which has been recently introduced as an independent

control to verify that the BST or SNO-RAC are indeed detecting an SNO moiety, and not

other Cys-based redox modifications [110, 116]. Pre-photolysis abrogated the GAPDH-

SNO signal, consistent with the established sensitivity of the S-NO bond to UV-

homolysis [75] (Figure 18A).

In an attempt to discover other endogenously S-nitrosylated proteins whose steady

state levels reflect iNOS, rather than denitrosylase activity, we subjected resting and

cytokine-stimulated macrophages to SNO-RAC with protein visualization by Coomassie

staining. We readily detected a single protein-SNO at ~ 22 kD, which was excised and

identified by MALDI-TOF MS as peroxiredoxin-1 (Prx1) (Figure 18B). To confirm these

findings, we subjected resting and cytokine-stimulated macrophages to SNO-RAC ±

ascorbate and pre-photolysis. As shown in Figure 18C, both the expression of Prx1 and

the degree of SNO-RAC pulldown are increased by cytokine stimulation. Further, the

omission of ascorbate completely abrogates the Prx1-SNO signal, and pre-photolysis also

attenuates the signal, thus verifying that Prx1 is indeed S-nitrosylated. These data

highlight the utility of combining pre-photolysis with the BST/SNO-RAC for SNO-

detection under conditions of altered protein expression, which might otherwise confound

Page 85: thesis mtf 1 - Duke University

70

the results. In HEK293 cells transfected with iNOS, Prx1-SNO was also readily detected

by SNO-RAC (Figure 18D), further validating that this protein is S-nitrosylated in vivo.

4.7 Detection of S-Acylation by a Modified SNO-RAC: “Acyl-RAC”

BST-like methodology has been adapted for the identification of S-acylation (e.g.

S-palmitoylated Cys residues) by substitution of ascorbate with hydroxylamine [154], a

nucleophile that readily cleaves thioesters to generate free thiol and hydroxamic acid

(Figure 19A). To expand the utility of SNO-RAC, we developed an analogous assay for

solid-phase capture of S-acylated proteins by substituting ascorbate with hydroxylamine

(dubbed “Acyl-RAC”). In addition, captured proteins were labeled “on-resin” using an

amine-reactive fluorophore (NHS-Cy5), and eluted proteins were visualized by

fluorescence after SDS-PAGE. Utilizing these two techniques, S-palmitoylated proteins

from bovine brain membranes (which are highly enriched for lipid-modified proteins)

were readily detected in a hydroxylamine-dependent manner (Figure 19B). Importantly,

this finding demonstrates both the versatility of resin-based methodology for the

detection of various Cys-based posttranslational modifications and the robust

compatibility of resin-immobilization with covalent labeling (e.g. Cy3/Cy5

fluorophores).

4.8 Mapping SNO-Sites by SNO-RAC and Label-Free LC-MS/MS

To assess proteome-wide S-nitrosylation, we subjected RAW264.7 cells to an

exogenous nitrosative stress (200 µM CysNO x 10 min) and subjected the lysates to

SNO-RAC. Following protein immobilization, resins were treated with trypsin, washed

Page 86: thesis mtf 1 - Duke University

71

and peptides eluted with reductant. This approach removes all peptides that are not

covalently bound to the resin, and therefore reduces a set of protein-SNOs to individual

peptides containing specific SNO-sites (i.e. Cys residues). As shown by the LC-MS/MS

chromatogram in Figure 20A, SNO-RAC isolated significantly more peptides following

treatment with CysNO. Shown in Figure 20B is the triply-charged precursor ion from a

single representative peptide, which was readily identified by MS/MS as the active site

Cys of the ubiquitin conjugating enzyme UbcH7 (Figures 20C and 20D), a major

intracellular ubiquitin ligase and cell cycle regulator [155, 156]. This active site Cys thiol

is responsible for ubiquitin transfer via formation of a glycyl thioester with ubiquitin, and

thus S-nitrosylation of this active site would likely inhibit its activity. The remaining 47

identified peptides are listed in Table 2. Importantly, 46 of the 47 detected peptides

contain Cys residues, which was not a constraint in our database search and thus re-

confirms the notion that SNO-RAC does indeed isolate SNO-sites. Although it was not

determined why a single non-Cys-bearing peptide was isolated (from heterogenous

ribonucleoprotein L), some peptides may exhibit affinity for the sepharose matrix. These

observed SNO-sites include previously reported targets of endogenous S-nitrosylation,

such as the active site Cys150 of GAPDH [27], as well as 5 SNO-sites reported by Hao et

al. from CysNO-treated rat cerebellum: Cys179/159 of Erk1/2, Cys295 and 349 of α-

tubulin, Cys369 of phosphoglycerate dehydrogenase, and Cys109 of poly-r(C) binding

protein [145].

Page 87: thesis mtf 1 - Duke University

72

4.9 Mapping SNO-Sites by SNO-RAC and Isotopically Encode LC-MS/MS

An increasingly popular approach in proteomics is to employ isotopic labeling to

allow comparison of control vs. test samples in a single MS experiment [157, 158].

Importantly, isotopic labeling minimizes experimental variation between individual

chromatographic separations. Since SNO-RAC is well suited to on-resin trypsinization

and covalent labeling, we sought to employ isotopic labeling for SNO proteome

characterization in E. coli subjected to an exogenous nitrosative stress. To this end,

lysates from untreated or CysNO-exposed E. coli were subjected to SNO-RAC and on-

resin proteolysis, after which the two samples were acetylated with H6- (control) or D6-

(CysNO-treated) acetic anhydride (these reagents react with all amino groups (N-termini

and lysine side chains), thus generating a mass difference of 3.03 Da for each reaction).

MG1655 E. coli were either untreated or incubated with 500 µM CysNO for 5 min

(representing a growth-inhibitory nitrosative stress), and each were subjected to SNO-

RAC. Following on-resin trypsinization, the untreated and CysNO-treated samples were

amino-acylated with H6- and D6-acetic anhydride, respectively, thus generating a mass

difference between the control and CysNO-treated samples. After elution and mixing of

the two samples, peptides were identified by LC-MS/MS. This approach readily revealed

44 previously unidentified specific sites of protein S-nitrosylation in E. coli (Table 3),

and all but one identified peptide contained a Cys residue. Importantly, all observed

precursor ion masses were within 0.05 Da of the calculated D3-acetylated peptide mass

and the lack of H3-acetylated peptides eliminates the possibility of nonspecific or false-

positive peptide identifications. To independently confirm the presence and absence of

D3- and H3-labeled peptides, respectively, the same samples were analyzed by MALDI-

Page 88: thesis mtf 1 - Duke University

73

TOF MS (Figure 21). Mass spectra from two representative SNO-sites are shown in

Figures 21B and 21C; in each case, the D3-labeled mass envelope is observed, indicating

these peptides were obtained only following CysNO treatment.

4.10 Discussion

The BST has become a widely used method for the detection and discovery of

protein-SNOs, and it has served to facilitate many studies probing S-nitrosylation in vivo.

While it is certainly an excellent technique—as evidenced by the greater than 200 studies

that have employed it—it does possess several drawbacks: it is highly operator-

dependent, labor intensive and low throughput. These disadvantages are mostly due to the

large number of steps required to perform a BST (i.e. multiple acetone precipitations,

biotin-HPDP removal, avidin pulldown). Here we demonstrate that direct resin-based

immobilization of protein-SNOs (SNO-RAC), in lieu of thiol biotinylation and avidin

pulldown, circumvents several of these hindrances. One of the advantages of SNO-RAC

is that it requires only a single acetone precipitation (to remove MMTS), which makes it

much easier to perform relative to the BST. Further, the direct covalent conjugation of

proteins to the resin via a stable disulfide allows for buffer exchanges or chemical

manipulations that would not normally be feasible with the BST. As demonstrated here,

immobilized proteins are readily trypsinized “on-resin”, which permits the conversion of

protein-SNOs to tryptic peptides containing individual SNO sites (i.e. peptides containing

the S-nitrosylated Cys residue of interest). These methodological advantages of SNO-

RAC over the BST offer a significant advance in the field of SNO proteomics.

Page 89: thesis mtf 1 - Duke University

74

The SNO-site identifications from mouse macrophages and E. coli reveal a total

of 91 specific sites of S-nitrosylation, the majority of which are novel identifications. The

E. coli peptide-SNO identifications listed in Table 3 reveal that numerous ribosomal

proteins are targets of S-nitrosylation, which suggests the possibility that nitrosative stress

exerts bacteriostatic/bacteriocidal effects via ribosomal perturbations, potentially similar

to the plethora of ribosome-directed antibiotics (e.g. macrolides, aminoglycosides) [159].

In the case of ribosomal protein S12 (gene rpsL), one of the identified SNO sites (Cys27)

was previously shown to be critical for maintaining translational efficiency and accuracy

[160]. Another novel SNO target, Cys298 of tryptophanase (gene tnaA), has been shown to

regulate substrate affinity and protein stability [161]. An additional SNO site with

potential relevance to the mechanisms of nitrosative stress was detected in MreB, an

ancient actin-like protein that has recently been shown to orchestrate chromosomal

segregation [162], replication and cytoskeletal organization [163, 164], and which has

become the target of numerous efforts for antibiotic development [165, 166]. The

identified site of S-nitrosylation within MreB (Cys113) is highly conserved throughout

bacterial MreB orthologs, consistent with the possibility that nitrosative stress exerts a

cytostatic/toxic effect via disruption of MreB-dependent cellular functions.

Notwithstanding the major advances in SNO detection provided by SNO-RAC vs.

the BST, particularly with respect to the simplicity and versatility of SNO-RAC, it should

be noted that the sensitivities of both assays are enhanced significantly by western

blotting for a specific protein-SNO. The same situation applies to the study of other

posttranslational modifications, such as protein O-phosphorylation, where antibodies to

specific phosphorylated sites often detect phosphorylation events that are undetectable on

Page 90: thesis mtf 1 - Duke University

75

the proteomic scale. Efforts to increase the degree of phosphorylation (e.g. phosphatase

inhibition with vanadate) are therefore often employed in proteomic studies. It remains to

be seen whether the inhibition of denitrosylation reactions will similarly facilitate the

proteomic mining of endogenous protein-SNOs. Collectively, our experiments

demonstrate that SNO-RAC is a highly versatile and simple methodology that should

greatly facilitate the study of SNO proteomes.

Page 91: thesis mtf 1 - Duke University

76

Figure 15. Schematic illustration demonstrating the enrichment and analysis of protein S-nitrosylation by resin-assisted capture (SNO-RAC). First, proteins thiols are methylthiolated (“blocked”) with MMTS; next, ascorbate is used to selectively convert protein SNOs to free thiols via a transnitrosation reaction; finally, nascent thiols are covalently “trapped” with a resin-bound 2- or 4-pyridyl disulfide (denoted as “X”). After washing, the resin-bound proteins are eluted with reductant and analyzed by SDS-PAGE and western blotting. Alternatively, proteins can be either “on-resin” labeled with amine-reactive fluorophores (e.g. Cy3/Cy5) or trypsinized to obtain Cys-containing peptides that contain sites of S-nitrosylation (SNO-sites). To employ isotopic labeling, the immobilized tryptic peptides can be incubated with amine-reactive isotopic labels (denoted as “L” and “H” for light and heavy isotopes, respectively), followed by elution and analysis by MALDI-TOF or LC-MS/MS.

Page 92: thesis mtf 1 - Duke University

77

Figure 16. SNO-RAC detects S-nitrosylation in vitro and maps SNO sites within purified GAPDH. Rabbit GAPDH (25 µM, 400 µg total) was treated with buffer or 500 µM CysNO, desalted to remove remaining CysNO and subjected to SNO-RAC. (A) Following elution with DTT, GAPDH-SNO was detected by Coomassie stain or western blotting. (B) Immobilized GAPDH-SNO was incubated with 0.5 µg sequencing grade trypsin at 37 °C for 12 h and protein digestion was assessed by western blotting. (C) Following elution of peptides generated by on-resin trypsinization of GAPDH-SNO, each Cys-containing peptide was identified by MALDI-TOF MS, demonstrating the facile detection of numerous S-nitrosylated GAPDH Cys residues. The asterisk indicates a peptide with a partial missed cleavage.

Page 93: thesis mtf 1 - Duke University

78

Figure 17. A side-by-side comparison of the BST and SNO-RAC with CysNO-treated HEK293 cells. Cultured HEK293 cells were either untreated or incubated with 500 µM CysNO for 10 min, extracts were harvested and subjected to either the BST or SNO-RAC. (A) Eluates from both experiments were analyzed by SDS-PAGE and Ag staining, which demonstrates improved sensitivity for high molecular weight protein-SNOs by SNO-RAC. These protein-SNOs obtained only with SNO-RAC (not with the BST) were gel-excised and identified by LC-MS/MS (Table 1 contains a detailed list). Avidin co-eluted in the BST is indicated by an asterisk. (B) A side-by-side comparison of the BST and SNO-RAC with CysNO-treated HEK293 cells reveals comparable sensitivities for representative lower mass protein-SNOs, such as protein tyrosine phosphatase 1B (PTP1B), caspase-3 and thioredoxin-1 (Trx1).

Page 94: thesis mtf 1 - Duke University

79

Figure 18. SNO-RAC can be employed to analyze and discover endogenous protein-SNOs. (A) RAW264.7 murine macrophages were left untreated or stimulated with 0.5 µg/ml LPS and 100 U/ml IFNγ to induce iNOS, lysed and subjected to SNO-RAC. Prior to blocking, some extracts were subjected to UV “pre-photolysis” to verify the specificity of SNO-RAC. As demonstrated, endogenously S-nitrosylated GAPDH is readily detected by SNO-RAC. (B) Macrophages were treated and subjected to SNO-RAC as in panel A, but the gel was visualized by Coomassie staining, which revealed a prominent band at ~ 22 kD following cytokine stimulation. This putative protein-SNO was excised and identified by MALDI-TOF MS as peroxiredoxin-1 (Prx1). (C) SNO-RAC followed by western blotting for Prx1 reveals that cytokine stimulation leads to increased pulldown of Prx1, which is dependent on the inclusion of ascorbate in the assay and attenuated by pre-photolysis, confirming the true identity of Prx1-SNO. (D) To verify the role of iNOS in the S-nitrosylation of Prx1, HEK293 cells were transfected for 18 h with plasmid ± iNOS cDNA, lysates were subjected to SNO-RAC and Prx1-SNO was assessed by western blotting.

Page 95: thesis mtf 1 - Duke University

80

Figure 19. SNO-RAC can be modified to detect S-acylated, rather than S-nitrosylated proteins. (A) hydroxylamine (NH2OH) is a nucleophile that specifically cleaves thioester groups into free thiol and hydroxamic acid. It is frequently employed to cleave a radiolabeled acyl group (e.g. 3H-palmitate) from a protein of interest to demonstrate protein S-palmitoylation. By replacing ascorbate with hydroxylamine, S-acylated/palmitoylated proteins are readily detected by resin-assisted capture (dubbed “Acyl-RAC”). (B) Bovine brain membranes (2 mg total), which are highly enriched in S-acylated proteins, were subjected to Acyl-RAC. Prior to elution, immobilized proteins were labeled with amine-reactive NHS-Cy5 and the gel was imaged by fluorescence, thus demonstrating the compatibility of on-resin covalent labeling with SNO-RAC/Acyl-RAC.

Page 96: thesis mtf 1 - Duke University

81

Figure 20. Compatibility of SNO-RAC with on-resin trypsinization and mass spectrometry-based SNO-site identification. Cultured RAW264.7 Macrophages were either untreated or incubated with 200 µM CysNO for 10 min, lysed and subjected to SNO-RAC. Following on-resin trypsinization to convert immobilized protein-SNOs to specific SNO-sites, peptides were eluted with DTT and analyzed by LC-MS/MS. (A) The reverse-phase chromatogram from triplicate analyses demonstrates a significantly increased abundance of peptide isolation following CysNO treatment. The red and blue traces correspond to control and CysNO-treated samples. (B) A representative precursor ion spectrum from a SNO-site, indicated by an asterisk in panel (A). (C) Following MS/MS fragmentation, this SNO-site was identified as the active site Cys53 of ubiquitin conjugating enzyme E2 L3 (UbcH7), implying that S-nitrosylation of this enzyme would inhibit its function as a ubiquitinylating enzyme. (D) A list of the observed b- and y-series ions from the MS/MS spectrum (in red and blue, respectively) demonstrates that SNO-RAC readily yields SNO-site identifications. Samples were analyzed using a Waters MSE on a QToF Premier Mass spectrometer. Test and control samples were analyzed three times each and data were processed using ProteinLynx Global Server (PLGS) v2.3 with the IdentityE search algorithm, using 15 ppm precursor and 30 ppm product ion tolerances. Peptides were considered as a confident identification if there were 8 or more product ions matched, and these are listed in Table 2.

Page 97: thesis mtf 1 - Duke University

82

Figure 21. MALDI-TOF analysis of E. coli SNO-sites identified by isotopically encoded SNO-RAC. A 50 ml culture of K-12 E. coli in M9 minimal media containing 0.2% glucose was grown to OD600 1.0, then left untreated or exposed for 5 min to 500 µM CysNO and cellular extracts were subject to SNO-RAC. Following on-resin trypsinization, the untreated and treated samples were acetylated with H6- and D6-acetic anhydride, respectively. The two resins were then combined, eluted with 10 mM DTT, concentrated by speedvac and directly analyzed by MALDI-TOF MS. (A) The full MALDI spectrum reveals a large number of single-isotopically labeled peptides, the identity of which (i.e. H3- vs. D3-labeled) was determined by LC-MS/MS (Table 3). (B) An expanded view of the SNO peptide containing Cys53 of Ribosomal protein L27 (RL27) demonstrates that only the D3-, and not the H3-labeled peptide was obtained. (C) An expanded view of the SNO peptide containing Cys298 of tryptophanase (TnaA) demonstrates that only the D3-, and not the H3-labeled peptide was obtained. These two SNO-sites are representative of the remaining SNO proteome. Since each acetylation occurs on primary amines, the potential H3-labeled monoisotopic (M+H)+ peptides for RL27 and TnaA would be ~ 6.06 and 3.03 amu smaller than the observed D3-labeled monoisotopic (M+H)+ peaks, respectively (i.e. peptides cleaved at C-terminal Arg or Lys contain 1 or 2 total amino groups, including the N-terminal amine). Note that peaks in these MALDI spectra are one proton larger than the predicted masses due to protonation and detection in positive-ion mode.

Page 98: thesis mtf 1 - Duke University

83

Table 1. Identifications of high molecular weight protein-SNOs by SNO-RAC coupled with SDS-PAGE and LC-MS/MS. HEK293 cells were treated as described in Fig 1 and subjected to SNO-RAC. Coomassie-stained protein bands >150 kD that were visualized only from SNO-RAC and not the BST were excised and analyzed by LC-MS/MS. Fifteen novel protein-SNOs were isolated. It is worth noting that this experiment yields whole protein identifications, not specific SNO-sites as obtained from “on-resin” trypsinization.

Protein Swissprot MW

(kDa) %

Coverage Unique

Peptides Protein Name Function/Phenotypes

PRKDC P78527 469 18.0 68 DNA-dependent protein kinase; catalytic subunit

Serine/threonine-protein kinase that acts as a molecular sensor for DNA damage

FLNA P21333 281 16.2 29 Filamin-A Anchors proteins to actin cytoskeleton

FLNB O75369 278 17.9 37 Filamin-B Anchors proteins to actin cytoskeleton

FAS P49327 273 11.1 24 Fatty acid synthase Synthesizes long-chain fatty acids from acetyl-CoA, malonyl-CoA and NADPH

PYR1 P27708 243 8.72 17 CAD protein Multifunctional enzyme in pyrimidine biosynthesis pathway

SC16A O15027 234 5.32 9 Protein transport protein Sec16A

Directs secretory cargo from ER to the Golgi apparatus

USP9X Q93008 290 1.73 4 Probable ubiquitin carboxyl-terminal hydrolase FAF-X

Removes ubiquitin from target proteins, thereby preventing their degradation

ANR17 O75179 274 2.23 5 Ankyrin repeat domain-containing protein 17

Function unknown; marker of hepatic differentiation; target of enterovirus 71

WNK1 Q9H4A3 251 2.23 4 Serine/threonine-protein kinase WNK1

Sensor of hypertonicity; controls Na+ and Cl- transport by inhibiting WNK4

LRBA P50851 319 1.36 3 Lipopolysaccharide-responsive and beige-like anchor protein

Guides intracellular vesicles to LPS-activated receptor complexes

CBP Q92793 265 1.88 3 CREB-binding protein Acetylates histones; binds to and enhances CREB transcriptional activity

FLNC Q14315 291 2.64 3 Filamin-C Anchors proteins to actin cytoskeleton; predominantly in muscle

AKA12 Q02952 191 1.35 2 A-kinase anchor protein 12 Anchors Protein Kinase A and C to actin cytoskeleton

MIA3 Q5JRA6 214 1.21 2 Melanoma inhibitory activity protein 3 precursor

Transmembrane protein; suppressor of malignant melanoma

Page 99: thesis mtf 1 - Duke University

84

Table 2. Identifications of specific SNO-sites in CysNO-treated RAW264.7 macrophages by SNO-RAC coupled with “on-resin” trypsinization and LC-MS/MS. Cultured RAW264.7 cells (4 x 15 cm dishes total) were incubated ± 200 µM CysNO for 10 min, collected and 6 mg each lysate was subjected to SNO-RAC with on-resin trypsinization. Following elution with 5 mM DTT, samples were concentrated by Speedvac and directly analyzed by LC-MS/MS. Triplicate MS analyses were performed for both control and CysNO-treated samples. The database search was performed with 15 ppm precursor and 30 ppm product ion mass tolerances, and identifications were considered confident if at least 8 product ions were observed.

Name Swiss-prot

Δ mass (ppm)

Fold Change Peptide Sequence Protein/Function

PP2Aα Q76MZ3 3.9 73.6 LNIISNLDCVNEVIGIR

Serine/threonine-protein phosphatase 2A; 65 kDa regulatory subunit A alpha isoform; general Ser/Thr phosphatase with broad substrate specificity

mAspAT P05202 1.6 14.8 TCGFDFSGALEDISK

Aspartate aminotransferase; mitochondrial precursor; facilitates cellular uptake of long-chain free fatty acids

ALDH2 P47738 3.3 1000.0 TFPTVNPSTGEVICQVAEGNKEDVDK

Aldehyde dehydrogenase; mitochondrial precursor; converts aldehydes to carboxylic acids

PRMT1 Q9JIF0 4.0 359.5 VEDLTFTSPFCLQVK Protein arginine N-methyltransferase 1; transcriptional and histone regulator

Annexin-A7 Q07076 3.4 1000.0 LGTDESCFNMILATR Annexin A7; promotes membrane

fusion and exocytosis

ATOX1 O08997 10.4 138.4 VCIDSEHSSDTLLATLNK Copper transport protein ATOX1; regulates Cu-ATPase

Cathepsin S O70370 0.9 32.5 NHCGIASYCSYPEI

Cathepsin S; lysosomal thiol protease, responsible for MHC class II processing

Cathepsin S O70370 4.2 1000.0 SGVYDDPSCTGNVNHGVLVV

GYGTLDGK

Cathepsin S; lysosomal thiol protease, responsible for MHC class II processing

Dynamin-2 P39054 3.6 115.2 LQDAFSSIGQSCHLDLPQIAVVGGQSAGK

Dynamin-2; GTPase responsible for mediating endocytosis

EF-1α P10126 2.6 1000.0 SGDAAIVDMVPGKPMCVESFSDYPPLGR

Elongation factor 1-alpha 1; promotes GTP-dependent binding of aminoacyl-tRNA to A-site of ribosomes

EF-1β O70251 0.1 14.9 SYIEGYVPSQADVAVFEAVSGPPPADLCHALR

Elongation factor 1-beta; stimulates exchange of GDP to GTP on EF-1α

EF-1δ P57776 3.7 97.2 SSILLDVKPWDDETDMAQLETCVR

Elongation factor 1-delta; ; stimulates exchange of GDP to GTP on EF-1α

FABP5 Q05816 12.8 264.0 MAAMAKPDCIITCDGNNITVK Epidermal fatty acid-binding protein; facilitates cellular differentiation

FABP5 Q05816 0.7 56.2 MIVECVMNNATCTR Epidermal fatty acid-binding protein; facilitates cellular differentiation

FPP synthetase Q920E5 2.1 1000.0 SLIEQCSAPLPPSIFMELANK

Farnesyl pyrophosphate synthetase; early enzyme in the biosynthesis of isoprenoids and cholesterol

Page 100: thesis mtf 1 - Duke University

85

Table 2, continued

GAPDH P16858 3.3 204.9 IVSNASCTTNCLAPLAK

Glyceraldehyde-3-phosphate dehydrogenase; glycolytic enzyme; effector of NO-dependent cell death via S-nitrosylation

GAPDH P16858 3.2 134.4 VPTPNVSVVDLTCR

Glyceraldehyde-3-phosphate dehydrogenase; glycolytic enzyme; effector of NO-dependent cell death via S-nitrosylation

hnRNP-K P61979 6.8 101.0 IIPTLEEGLQLPSPTATSQLPLESDAVECLNYQHYK

Heterogeneous nuclear ribonucleoprotein K; binds pre-mRNA and facilitates splicing

hnRNP-L Q8R081 10.9 72.9 NGVQAMVEFDSVQSAQR Heterogeneous nuclear ribonucleoprotein L; binds pre-mRNA and facilitates splicing

eIF-5A-2 Q8BGY2 11.1 331.8 YEDICPSTHNMDVPNIK Eukaryotic translation initiation factor 5A-2; promotes ribosomal peptide bond formation

JMJD1B Q6ZPY7 3.6 112.6 EVKEMAMGLNVLDPHTSHSWLCDGR

JmjC domain-containing histone demethylation protein 2B; demethylates histone H3 at Lysine 9

MAPK3/1 P63085 10.0 435.0 DLKPSNLLLNTTCDLK Mitogen-activated protein kinase 3/1; ERK1/2; stress- and growth-responsive kinase

p67-Phox O70145 2.5 185.8 GLVPCNYLEPVELR

Neutrophil cytosol factor 2; Nox2 p67 subunit; promotes respiratory burst and superoxide production; defective in chronic granulomatous disease

α-CP1 P60335 2.3 48.7 LVVPATQCGSLIGK Poly(rC)-binding protein 1; binds single stranded poly-cytosine DNA

α-CP2 Q61990 0.4 265.2 AITIAGIPQSIIECVK Poly(rC)-binding protein 2; binds single stranded poly-cytosine DNA

α-CP2 Q61990 2.3 360.0 LVVPASQCGSLIGK Poly(rC)-binding protein 2; binds single stranded poly-cytosine DNA

PCNA P17918 12.7 116.7 DLSHIGDAVVISCAK Proliferating cell nuclear antigen; mediates DNA replication; senses DNA damage

PCNA P17918 8.8 54.0 LMDLDVEQLGIPEQEYSCVIK Proliferating cell nuclear antigen; mediates DNA replication; senses DNA damage

PEBP-1 P70296 0.8 132.6 YVWLVYEQEQPLSCDEPILSNK

Phosphatidylethanolamine-binding protein 1; potentiates cholinergic neuron function; inhibits Raf-1 kinase

PGAM-1 Q9DBJ1 10.0 1000.0 YADLTEDQLPSCESLKDTIAR Phosphoglycerate mutase 1; glycolytic enzyme

Plastin-2 Q61233 2.2 93.4 VDTDGNGYISCNELNDLFK Plastin-2; binds to actin; component of cytoskeleton

RNH1 Q91VI7 5.1 1000.0 ELDLSNNCMGGPGVLQLLESLK

Ribonuclease inhibitor; forms stoichiometric complex with Ribonuclease H

RNH1 Q91VI7 6.9 1000.0 TNELGDGGVGLVLQGLQNPTCK

Ribonuclease inhibitor; forms stoichiometric complex with Ribonuclease H

Rplp0 P14869 6.2 1000.0 AGAIAPCEVTVPAQNTGLGPEK 60S acidic ribosomal protein P0

L10E P47955 2.2 8.6 ALANVNIGSLICNVGAGGPAPAAGAAPAGGAAPSTAAAPAEEK 60S acidic ribosomal protein P1

Page 101: thesis mtf 1 - Duke University

86

Table 2, continued Phospho-

protein NP33

P62754 0.1 944.2 LNISFPATGCQK 40S ribosomal protein S6

3-PGDH Q61753 1.7 54.8 NAGTCLSPAVIVGLLR D-3-phosphoglycerate dehydrogenase; enzyme responsible for serine biosynthesis

STI1 Q60864 1.3 1000.0 ALSAGNIDDALQCYSEAIK Stress-induced-phosphoprotein 1; mediates association of molecular chaperones HSC70 and HSP90

AsnRS Q8BP47 7.3 1000.0 GYCEVTTPTLVQTQVEGGATLFK Asparaginyl-tRNA synthetase

α-Tubulin 1B P68369 0.8 1000.0 AYHEQLSVAEITNACFEPANQ

MVK Tubulin alpha-1B chain; microtubule network

α-Tubulin 1B P05213 1.8 1000.0 SIQFVDWCPTGFK Tubulin alpha-1B chain; microtubule

network

α-Tubulin 1A P68369 2.9 264.7 TIQFVDWCPTGFK Tubulin alpha-1A chain; microtubule

network

TCP-1α P11983 6.5 108.9 YINENLIINTDELGRDCLINAAK T-complex protein 1 subunit alpha B; molecular chaperone for protein folding

TCP-1γ P80318 1.1 130.6 IPGGIIEDSCVLR T-complex protein 1 subunit gamma; molecular chaperone for protein folding

TCP-1ζ P80317 9.7 1000.0 NAIDDGCVVPGAGAVEVALAEALIK

T-complex protein 1 subunit zeta; molecular chaperone for protein folding

HSP75 Q9CQN1 6.8 54.1 LVCEGQVLPEMEIHLQTDAK Heat shock protein 75 kDa, mitochondrial; molecular chaperone

UbcH7 P68037 11.7 49.4 GQVCLPVISAENWKPATK Ubiquitin-conjugating enzyme E2 L3; transfers ubiquitin to target proteins; mediates p53 degradation

Page 102: thesis mtf 1 - Duke University

87

Table 3. Identifications of specific SNO-sites in CysNO-treated E. coli by isotopically encoded SNO-RAC coupled with and LC-MS/MS. A 50 ml culture of K-12 E. coli in M9 minimal media containing 0.2% glucose was grown to OD600 1.0, then left untreated or exposed for 2 min to 500 µM CysNO and cellular extracts were subject to SNO-RAC. Following on-resin trypsinization, the untreated and treated samples were acetylated with H6- and D6-acetic anhydride, respectively. The two resins were then combined, eluted with 10 mM DTT, and eluates were concentrated by speedvac and directly analyzed by LC-MS/MS. D3-acetylated peptides with Mascot scores sufficient to generate p values < 0.01, 0.05 and 0.10 are indicated. As expected, H3-acetylated peptides were absent, demonstrating that the identified peptides were derived from the CysNO-treated sample. One D3-labeled, non-Cys-bearing peptide was consistently identified (derived from the enzyme tryptophanase, from which a Cys-containing peptide was also detected), likely due to an inadvertent affinity of this peptide for the sepharose matrix.

Protein Swissprot Peptide MW

Δ mass (ppm)

Mascot Score P-value Peptide Sequence Function/Class

AK1H P00561 866.4438 12.7 36 0.01 VCGVANSK

bifunctional aspartokinase/homoserine dehydrogenase I (Lys, Met and Thr biosynthesis)

AK1H P00561 1775.0174 1.4 32 0.01 TITPIAQFQIPCLIK

bifunctional aspartokinase/homoserine dehydrogenase I (Lys, Met and Thr biosynthesis)

AROF P00888 1868.8924 0.5 40 0.01 GGKAPNYSPADVAQCEK

phospho-2-dehydro-3-deoxyheptonate aldolase, Tyr-sensitive (aromatic amino acid biosynthesis)

AROF P00888 1581.7251 2.3 32 0.01 APNYSPADVAQCEK

phospho-2-dehydro-3-deoxyheptonate aldolase, Tyr-sensitive (aromatic amino acid biosynthesis)

DAPA P0A6L2 915.3771 12.1 46 0.01 DMAQMCK dihydrodipicolinate synthase (Lys biosynthesis)

DCUP P29680 1259.5797 1.1 31 0.01 AQAGDFMSLCK uroporphyrinogen decarboxylase (porphyrin biosynthesis)

DEOB P0A6K6 1543.8266 1.4 33 0.01 IADIYANCGITKK phosphopentomutase (deoxynucleotide metabolism)

DHSB P07014 1959.0809 5.2 30 0.01 NGLACITPISALNQPGKK

succinate dehydrogenase iron-sulfur subunit (TCA cycle)

DLDH P0A9P0 1271.5644 1.5 24 0.05 AIASDCADGMTK

dihydrolipoyl dehydrogenase (E3 component of pyruvate and 2-oxoglutarate dehydrogenases)

G6PD P0AC53 1824.9562 3.5 54 0.01 AVTQTAQACDLVIFGAK

glucose-6-phosphate 1-dehydrogenase (pentose phosphate shunt, NADPH biosynthesis)

GATY P37192 1320.6767 0.1 52 0.01 VKEVVDFCHR tagatose-1,6-bisphosphate aldolase (galacticol metabolism)

Page 103: thesis mtf 1 - Duke University

88

Table 3, continued

GATZ P37191 1798.8421 3.1 37 0.01 AAVLCFAAESVATDCQR

tagatose 6-phosphate kinase (galacticol metabolism)

GLPB P13033 712.4139 9.7 29 0.05 KVTCK

glycerol-3-phosphate dehydrogenase subunit B (lipid and glycerophospholipid metabolism)

GRCA P68066 622.3631 12.7 36 0.01 CIVAK

glycyl radical cofactor (protects pyruvate-formate lyase from oxidative inactivation)

ILVB P08142 1540.7905 0.3 42 0.01 AIGKTEQFCPNAK

acetolactate synthase isozyme 1 large subunit (brached chain amino acid biosynthesis)

ILVC P05793 2046.0985 1.5 145 0.01 KVVIVGCGAQGLNQGLNMR

ketol-acid reductoisomerase (brached chain amino acid biosynthesis)

MREB P0A9X4 1669.9377 4.3 37 0.01 VLVCVPVGATQVERR

rod shape-determining protein (“actin-like” cytoskeletal protein)

NAPA P33937 1484.7121 12.5 28 0.05 VVACQGDPDAPVNR

periplasmic nitrate reductase (nitrate assimilation)

NFSA P17117 1414.714 3.5 57 0.01 MTPTIELICGHR oxygen-insensitive NADPH nitroreductase

NRDD P28903 1522.7715 8.9 32 0.01 VAPILYMEGACGVR

anaerobic ribonucleoside-triphosphate reductase (deoxyribonucleotide biosynthesis)

NUOB P0AFC7 1572.7799 5.8 38 0.01 QADLMVVAGTCFTK

NADH-quinone oxidoreductase subunit B (complex I of aerobic respiration)

PFLB P09373 856.3941 13.5 38 0.01 MIEGSCK formate acetyltransferase 1 (nonoxidative glucose metabolism)

PT1 P08839 1417.7427 2.7 115 0.01 AVAEACGSQAVIVR

phosphoenolpyruvate-protein phosphotransferase (phosphoenolpyruvate utilization)

RL2 P60422 1376.7796 0.7 52 0.01 AVVKCKPTSPGR 50S ribosomal protein L2

RL2 P60422 934.4813 0.4 45 0.01 CKPTSPGR 50S ribosomal protein L2

RL2 P60422 909.4496 17.8 27 0.05 KVEADCR 50S ribosomal protein L2

RL17 P0AG44 925.5326 0.8 21 0.1 ILKCGFR 50S ribosomal protein L17

RL27 P0A7L8 1989.975 6.7 45 0.01 FHAGANVGCGRDHTLFAK 50S ribosomal protein L27

RL28 P0A7M2 823.438 13.7 32 0.01 VCQVTGK 50S ribosomal protein L28

RL31 P0A7M9 1576.7674 1.8 50 0.01 STVGHDLNLDVCSK 50S ribosomal protein L31

Page 104: thesis mtf 1 - Duke University

89

Table 3, continued

RL31 P0A7M9 1723.7374 4.0 27 0.05 YEEITASCSCGNVMK 50S ribosomal protein L31

RL35 P0A7Q1 1910.0164 2.2 46 0.01 AMVSKGDLGLVIACLPYA 50S ribosomal protein L35

RPOA P0A7Z4 1408.71 0.1 78 0.01 LLVDACYSPVER DNA-directed RNA polymerase subunit alpha (RNA transcription)

RS12 P0A7S3 1501.7829 0.7 67 0.01 SNVPALEACPQKR 30S ribosomal protein S12

RS4 P0A7V8 1483.7472 1.3 44 0.01 CKIEQAPGQHGAR 30S ribosomal protein S4

SUCD P0AGE9 1704.8156 3.5 23 0.1 MIGPNCPGVITPGECK

succinyl-CoA ligase [ADP-forming] subunit alpha (TCA cycle)

TNAA P0A853 2995.4986 4.6 91 0.01 SYYALAESVKNIFGYQYTIPTHQGR

tryptophanase (tryptophan breakdown to pyruvate)

TNAA P0A853 2358.1064 3.0 23 0.05 TLCVVQEGFPTYGGLEGGAMER

tryptophanase (tryptophan breakdown to pyruvate)

TPX P0A862 3082.5916 4.1 114 0.01 KFNQLATEIDNTVVLCISADLPFAQSR

periplasmic thiol peroxidase (protection from oxidative stress)

TYPA P32132 1526.796 0.9 25 0.05 SNDLTVNCLTGKK

GTP-binding protein typA/BipA (virulence factor that regulates epithelial adhesion)

YDFZ P64463 1411.7215 4.0 30 0.05 GKTVVVEGCEEK putative selenoprotein (function unknown)

YEAO P76243 825.3995 17.1 36 0.01 MNIQCK uncharacterized protein

YHES P63389 1723.9124 3.8 27 0.05 KAELTACLQQQASAK

uncharacterized ABC transporter ATP-binding protein

YKGE P77252 2004.0079 2.0 50 0.01 MNVNFFVTCIGDALKSR uncharacterized protein

Page 105: thesis mtf 1 - Duke University

90

Chapter 5. Regulated Protein Denitrosylation by the Thioredoxin System

My primary reason for joining our laboratory was to investigate the mechanisms

by which S-nitrosylation is regulated. After my first few months in the lab, it became

more apparent that denitrosylation was a major determinant of overall protein S-

nitrosylation (akin to phosphorylation). Further, the topic of denitrosylation is

experimentally tractable by starting with a known S-nitrosylated protein and then

following its denitrosylation either in vitro or in cells.

Around this time, a senior post-doc in our lab, Moran Benhar, had observed that

the thioredoxin system catalyzes denitrosylation in intact cells. As bait, he employed S-

nitrosylation caspase-3, an exemplar of stimulus-induced denitrosylation and apoptosis.

Given our shared interest, he kindly allowed me to join in on his work. Specifically, I

performed assays demonstrating that overexpression of thioredoxin reductase 1 increases

the rate of protein denitrosylation in HEK293 cells. I also performed DNA fragmentation

assays with Fas-stimulated 10C9 B-lymphocytes. Visualizing the precisely ordered

breakdown of an entire mammalian genome, at the molecular level, is exciting and

humbling!

Benhar, M., Forrester, M.T., Hess, D.T. & Stamler, J.S. Regulated protein denitrosylation by cytosolic and mitochondrial thioredoxins. Science 320, 1050-1054 (2008).

Page 106: thesis mtf 1 - Duke University

91

5.1 Summary

Nitric oxide acts in cellular signal transduction through stimulus-coupled S-

nitrosylation of cysteine residues. The mechanisms that might subserve protein

denitrosylation in cellular signaling remain uncharacterized. Our search for denitrosylase

activities focused on caspase-3, an exemplar of stimulus-dependent denitrosylation, and

identified thioredoxin in a biochemical screen. In resting human lymphocytes,

thioredoxin-1 actively denitrosylated cytosolic caspase-3 and thereby maintained a low

steady-state amount of S-nitrosylation. Upon stimulation of Fas, thioredoxin-2 mediated

denitrosylation of mitochondria-associated caspase-3, required for its activation, and

promoted apoptosis. Inhibition of thioredoxins enabled identification of additional

substrates subject to endogenous S-nitrosylation. Thus, specific enzymatic mechanisms

may regulate basal and stimulus-induced denitrosylation in mammalian cells.

Page 107: thesis mtf 1 - Duke University

92

5.2 Introduction

Cellular functions of nitric oxide (NO) are carried out in part through S-

nitrosylation of Cys residues within a broad functional spectrum of proteins [21]. Like

protein phosphorylation, S-nitrosylation thus mediates or modulates transduction of

myriad cellular signals. Although the NO synthases (NOSs) and S-nitrosoglutathione

reductase (GSNOR) govern S-nitrosylation by NO or S-nitrosothiols (SNOs) [21], little is

known about the nature of or even necessity for enzymatic mechanisms that may directly

remove NO groups from Cys thiols.

Some caspases, members of a family of proteases that mediate apoptosis, are

subject to inhibitory S-nitrosylation, and stimulation of death receptors of the tumor

necrosis factor superfamily results in caspase denitrosylation [24, 25, 167]. In particular,

a subpopulation of caspase-3 (a major executioner caspase) that is associated with

mitochondria is constitutively S-nitrosylated at the active site Cys, and engagement of the

Fas receptor promotes denitrosylation [25, 107]. However, the role of denitrosylation in

transduction of the apoptotic signal has not been fully elucidated and the molecular

mechanism of denitrosylation is unknown.

5.3 Materials and Methods

Materials – Propylamine NONOate (PAPA-NONOate) and NG-monomethyl-L-

arginine (L-NMMA) were from Cayman Chemical. Stock solutions of PAPA-NONOate

were prepared in 10 mM NaOH. S-nitroso-cysteine (CysNO) was synthesized from L-

cysteine using acidified nitrite. Auranofin was from BioMol. Mouse monoclonal

Page 108: thesis mtf 1 - Duke University

93

antibodies to human caspase-3, thioredoxin 1 (Trx1) and protein tyrosine phosphatase 1B

(PTP1B) were from BD Biosciences. Polyclonal antibody to iNOS and mouse

monoclonal antibody to GAPDH were from Chemicon. Polyclonal antibodies to Trx2,

Trx1 reductase (TrxR1) and TrxR2 were from Abcam. Rabbit antiserum to human

thioredoxin-like protein 14 (TRP14) was generously provided by Dr. S. G. Rhee.

Polyclonal antibodies to human caspase-3 and caspase-9 and monoclonal antibody to

human caspase-8 were from Cell Signaling Technology. Monoclonal anti-Fas antibody

(human activating, clone CH11) was from Upstate Biotechnology. The human pRSC-

caspase-3 and pET23b-caspase-3 plasmids were generously provided by Dr. E. S.

Alnemri and Dr. G. S. Salvesen, respectively. Human TrxR1 (GenBank accession

number BC018122, IMAGE clone ID 3883490) in the vector pCMV-SPORT6 was from

American Type Culture Collection (ATCC). pEBG-GST-Trx, pEBG-GST-Trx(C35S)

and pET28a-Trx were generously provided by Dr. K. Kwon. Trx(C32S) in vector

pET28a was generated by site-directed mutagenesis using the QuickChange Kit

(Stratagene). Recombinant rat TrxR1 was from American Diagnostica. Recombinant

human Trx1 from E. coli or human T-cells, and all other reagents, were from Sigma

unless indicated.

Cell culture – HeLa, RAW264.7 and human embryonic kidney (HEK-293) cells

were maintained in DMEM supplemented with 10% FBS and 1% penicillin/streptomycin

at 37°C under 5% CO2. Jurkat (human T leukemia cell line) and 10C9 (human Burkitt's

lymphoma B cell line) cells were maintained in RPMI medium 1640 supplemented with

10% FBS and 1% penicillin/streptomycin at 37°C under 5% CO2.

Page 109: thesis mtf 1 - Duke University

94

Overexpression and purification of recombinant proteins – Human caspase-3 was

expressed in E. coli strain BL21(DE3). The His6-tagged protein was purified by affinity

chromatography on nickel-nitrilotriacetic acid-agarose (Qiagen) as described [168].

Briefly, bacterial lysates were prepared in buffer (50 mM Tris, pH 8.0, and 100 mM

NaCl) and incubated with nickel-nitrilotriacetic acid-agarose for 1 hour. Caspase-3 was

eluted with an imidazole linear gradient from 0-200 mM. Fractions containing caspase-3

were pooled and concentrated by ultrafiltration. Because caspase-3 oxidizes

spontaneously during purification, the enzyme was reduced with 100 mM DTT for 30

min at room temperature before use. Excess DTT was removed by Sephadex G-25

(Amersham) chromatography. GST-tagged Trx and His-tagged Trx were expressed in E.

coli and purified on GSH-agarose or nickel columns, respectively.

S-Nitrosylation of caspase-3 and reactivation assay – Reduced caspase-3 (10 µM

in buffer H: 50 mM Hepes, pH 7.5, 100 mM NaCl and 0.1% 3-[(3-

Cholamidopropyl)dimethylammonio]-1-propanesulfonate) was S-nitrosylated (SNO-

caspase-3) by exposure to 2 mM PAPA-NONOate (half-life ~ 77 min; 2 moles NO per

mole PAPA-NONOate) for 1 hour at room temperature followed by buffer exchange

through Sephadex G-25 equilibrated with buffer H. Under these conditions, we observed

a stoichiometry of ~ 0.9 mol SNO per mol caspase, as measured by Hg-coupled

photolysis chemiluminescence [25]. The flowthrough, which contains SNO-caspase-3,

was then diluted 1:4 with buffer H followed by incubation in 96-well HisGrab plates

(Pierce) for 1 hour at room temperature. Unbound material was removed by 3 washes

with buffer H, and immobilized SNO-caspase-3 (~ 100 nM) was incubated for 30 min

with cell extracts (diluted to 1 mg protein per ml in buffer H that contained 0.1 mM

Page 110: thesis mtf 1 - Duke University

95

diethylenetriaminepentaacetic acid) or with recombinant proteins. After 3 washes with

buffer H, caspase activity was measured with the EnzCheck kit (Molecular Probes) using

N-acetyl-(Asp-Glu-Val-Asp)-7-amido-4-methylcoumarin (Z-DEVD-AMC) and

evaluated at 340/450 nm (excitation/emission). In selected experiments, caspase was

eluted from the plate and SNO content was determined by Hg-coupled photolysis-

chemiluminescence [25].

Chromatographic purification of cellular denitrosylating activity – Jurkat cells (5

L culture) were collected, washed once in phosphate-buffered saline (PBS), resuspended

in isotonic buffer (10 mM Tris pH, 7.6, 210 mM mannitol, 70 mM sucrose) supplemented

with 1 x protease inhibitor mixture (Roche), and homogenized with 30-35 strokes of a

Dounce homogenizer (Wheaton). Nuclei and unbroken cells were removed by

centrifugation at 1000g for 10 min. The postnuclear supernatant was centrifuged at

20,000g for 30 min, and the resultant supernatant was loaded onto a Mono Q HR 10/10

column pre-equilibrated with 10 mM Tris, pH 7.5, and developed with a linear salt

gradient from 0 to 1 M NaCl. Active fractions were pooled, concentrated and loaded onto

a hydroxyapatite column. The column was washed with 25 mM sodium phosphate and

developed with a 25 mM – 400 mM sodium phosphate gradient (pH 7.6). After this step,

denitrosylating activity (>90%) was present in the unbound material, and was dependent

on the addition of eluted material (fraction II). Fraction II had negligible activity, but

when added in limiting amounts (~10% of unbound protein) reconstituted full activity of

the unbound fraction and of material purified during subsequent steps. The unbound

material from the last step was then loaded onto a Phenyl-Sepharose column, which was

eluted with a gradient of decreasing ammonium sulphate concentration (0.75-0 M). The

Page 111: thesis mtf 1 - Duke University

96

concentrated active fractions were then loaded onto a Supredex-200 column pre-

equilibrated with 100 mM Tris, pH 7.5, at a flow rate of 0.5 ml/min. One ml fractions

were collected, and SNO-caspase-3 denitrosylating activity was assessed using the

caspase reactivation assay. Proteins (5 mg) from the most active fraction (designated,

fraction I) were separated by SDS-PAGE (8-16% gradient) and visualized using Brilliant

Blue-G Colloidal Stain (Sigma). All protein bands were identified using an ABI 4700

matrix-assisted laser desorption/ionization time-of-flight/time-of-flight (MALDI-

TOF/TOF) tandem mass spectrometer (MS), internally calibrated with trypsin

autoproteolysis peaks. The MS spectrum was searched against the National Center for

Biotechnology Information (NCBI) database, using the on-line version of Protein

Prospector (http://prospector.ucsf.edu/). MS analysis was carried out at Michael Hooker

Proteomics and Mass Spectrometry Facility, University of North Carolina.

Detection of protein S-nitrosylation with biotin-switch and DAF-2 assays –

Detection of endogenously S-nitrosylated proteins was performed via the biotin-switch

method [50] with some modifications [116], and as follows. Whole cell lysates were

prepared with lysis buffer containing N-ethylmaleimide (50 mM Hepes, pH 7.5, 1% NP-

40, 150 mM NaCl, 0.1 mM EDTA, 10 mM NEM (Pierce), 1 mM phenylmethanesulfonyl

fluoride (PMSF) and protease inhibitor cocktail (Roche)). Fractions enriched for

mitochondria were prepared as described previously [107]. Lysates were adjusted to 1%

SDS and incubated at room temperature for 30 min in the dark to block free thiols. To

remove excess NEM, proteins were precipitated with 3 volumes of acetone at –20 °C for

30 min. The proteins were recovered by centrifugation at 5,000g for 5 min, and the

pellets were then resuspended in 240 µl of HENS buffer (250 mM Hepes, 1 mM EDTA,

Page 112: thesis mtf 1 - Duke University

97

0.1 mM neocuproine, pH 7.7, 1% SDS). After addition of sodium ascorbate (20 mM) and

biotin-HPDP (0.5 mM; Pierce), biotinylation was carried out for 1 hour at room

temperature. After removal of excess biotin-HPDP by acetone precipitation, biotinylated

proteins were pulled down overnight with streptavidin-agarose and analyzed by Western

blotting using West Femto substrate (Pierce). SNO-PTP1B was detected similarly, with

the blocking step performed at 50°C. Detection of SNO-GAPDH and UV exposures (to

validate SNO identification) were performed as previously described [116]. Typically, for

SNO quantitation, ascorbate-dependent signals were measured by densitometry and

normalized to total protein.

The DAF-2 assay was adapted from [169]. In brief, cell extracts (50 µ l) were

adjusted to 100 µ l with PBS containing DAF-2 (50 µ M) and HgCl2 (1 mM) and

incubated for 1 hour at room temperature, to yield the highly fluorescent

triazolofluorescein (DAF-2T). Fluorescence was monitored using excitation and emission

wavelengths of 485 and 520 nm, respectively.

Determination of soluble thiols – Acid-soluble (low molecular weight) thiols were

measured spectrophotometrically after reaction with 5,5’-dithio-bis(2-nitrobenzoic acid)

(DTNB, Ellman’s reagent). An equal volume of 10% metaphosphoric acid was added to

cell extracts, the precipitate was removed by centrifugation, and the supernatant was

neutralized with triethanolamine. Thiols were then measured by addition of an equal

volume of DTNB solution (1 mM final concentration). The absorbance of the Ellman’s

reagent adduct was measured at 405 nm. Values were derived by comparison with GSH

standards and were normalized to protein concentration in the extract.

Page 113: thesis mtf 1 - Duke University

98

Cell transfections – HEK and HEK-nNOS cells (seeded on polylysine-coated

culture dishes) were transiently transfected using Lipofectamine 2000 (Invitrogen)

following the manufacturer's instructions. Briefly, sub-confluent cells grown in 10 cm

dishes were transfected with 12 mg total DNA and 30 ml Lipofectamine 2000 in 3 ml of

Opti-MEM reduced-serum medium (Gibco). The plasmid DNA ratio of Trx-1:caspase-3

or TrxR1:caspase-3 was 2:1. Approximately 80% of HEK cells were successfully

transfected under these conditions as determined by a control transfection using a GFP

construct. When TrxR1 was overexpressed, the culture media was supplemented with 1

µM sodium selenite as described [170].

RNA interference and immunodepletion – Two small interfering RNAs (siRNAs)

(designated oligo 1 and 2) that target human Trx1 were based on nucleotides 258-276 and

149-167 relative to the translation start site; siRNA that targets human TRP14 was based

on nucleotides 189-209; siRNA that targets human TrxR2 were based on nucleotides

536-554 (oligo 1) and 1160-1178 (oligo 2); siRNA that targets human TrxR1 was based

on the smart pool approach (Dharmacon), and a second, individual oligo was based on

nucleotides 303-321; siRNA that targets human iNOS was based on nucleotides 302-320.

siRNAs were introduced into HeLa or HEK cells as described [171]. In brief, the cells

were harvested following exposure to trypsin, diluted with fresh antibiotic-free medium

and transferred to 6-well plates. After incubation for 24 hours, the cells were transfected

with siRNA for 3 days employing Oligofectamine (Invitrogen). The siCONTROL non-

targeting siRNA #1 (Dharmacon) served as a negative control for siRNA activity. A

mixture of 12 µ l of Opti-MEM reduced-serum medium and 3 µ l of Oligofectamine was

incubated for 10 min at room temperature and was then combined with 10 µl of siRNA

Page 114: thesis mtf 1 - Duke University

99

(20 µM) in 175 µ l of Opti-MEM. The resulting mixture (200 µ l) was incubated for 20

min at room temperature to allow complex formation and then overlaid onto each well of

cells to yield a final volume of 1.5 ml/well. siRNA oligonucleotides were introduced into

10C9 cells by square wave electroporation. Briefly, annealed siRNA duplexes (5 µM)

were mixed with cells resuspended in Opti-MEM medium, and the mixture was pulsed

once (2.5 kV/cm; 100 msec) using the Gene Pulser Xcell system (Bio-Rad). Three days

after electroporation, protein knockdown was determined by Western blotting.

For immunodepletion, antibodies (4 µg) were bound to protein A-agarose beads

(Amersham; 20 µl bead volume) in lysis buffer (50 mM Hepes, pH 7.5, 1% NP-40, 150

mM NaCl, 0.1 mM EDTA, 1 mM PMSF and protease inhibitor cocktail (Roche)),

brought to a total volume of 250 µl and incubated at 4 °C for 1 hour. Antibody-bound

beads were washed 3 times with buffer H to remove unbound antibodies. Freshly

prepared HeLa cell extract (50 µ l) was added to the beads and incubated with constant

rotation overnight at 4 °C. Beads were pelleted at 1000g for 2 min. Supernatants were

removed and used for in vitro assays and Western blotting.

GST pull-down assay – HEK cells were co-transfected with 4 µ g of pRSC-

caspase-3 plasmid and with 8 µg of empty vector, pEBG-GST-Trx1 or pEBG-GST-

Trx1(C35S). After 24 hours, cells were left untreated or treated with CysNO. Cell lysates

were then prepared and incubated with GSH-agarose beads overnight at 4°C. Beads were

washed 4 times in lysis buffer (with NaCl adjusted to 0.5M NaCl) and boiled in Laemmli

sample buffer, and eluted proteins were analyzed by Western blotting.

Affinity labeling of active caspases – Untreated or Fas-stimulated 10C9 cells were

lysed in 300 ml of lysis buffer that contained 10 µ M biotin-Val-Ala-Asp(OMe)

Page 115: thesis mtf 1 - Duke University

100

fluoromethyl ketone (bVAD-FMK). Lysates (1 mg protein) were incubated for 1 hour at

room temperature to allow labeling of caspase active sites. Biotinylated proteins were

captured by overnight incubation at 4°C with 30 µl streptavidin-agarose. Beads were

washed 3 times in lysis buffer containing 0.5 M NaCl, and biotinylated proteins were

eluted by incubation in 30 µ l Laemmli sample buffer at 95°C for 10 min. Extracts and

eluted, biotinylated proteins were analyzed by Western blotting with anti-caspase-3

antibodies.

Assessment of apoptosis by DNA fragmentation – Apoptotic DNA fragmentation

was assessed 6 hours after cell treatment with anti-Fas CH11 antibody. The DNA

fragmentation assay was performed as described [172] with minor modifications. In brief,

2-4 x 106 cells were lysed in 200 µl of lysis buffer (10 mM Tris, 10 mM EDTA, 0.5%

Triton X-100, pH 8.0) and centrifuged at 15,000g for 15 min. To 180 µl of the

supernatant was added 2 µg RNAse A for 2 hours at 37°C, followed by 2 µg proteinase K

for 1 hour at 37°C. Isopropanol (180 µl) was added and DNA was precipitated for 12-18

hours at -20°C. Following centrifugation at 10,000g for 5 min, DNA pellets were washed

with 70% ethanol and resuspended in 50 µl buffer (10 mM Tris, 1 mM EDTA, pH 8.0).

Aliquots of the re-solubilized DNA (30 µl) were analyzed by electrophoresis through a

1.2% agarose gel and visualized by ethidium bromide staining.

Statistical Analysis – Paired data were evaluated by Student's t-test. A one-way

ANOVA with the Tukey post-hoc test was used for multiple comparisons. A value of P <

0.05 was considered statistically significant. Results are shown as mean ± SEM.

Page 116: thesis mtf 1 - Duke University

101

5.4 Identification of Trx1 as a denitrosylating enzyme

To assess protein denitrosylation, we developed an assay based on the reactivation

of caspase that results from denitrosylation of the active site Cys (Figure 22A).

Recombinant caspase-3 was S-nitrosylated and the inhibited enzyme immobilized on

nickel-coated plates. Bound SNO-caspase-3 was incubated with cellular extracts, and

caspase-3 activity was then measured with a fluorogenic substrate. Addition of Jurkat cell

cytosolic extract rapidly restored the activity of caspase-3 (Figure 23A). Loss of the SNO

moiety was verified by Hg-coupled photolysis-chemiluminescence, which measures the

amount of NO displaced from Cys thiol by UV light [25] (Figure 23A). This cytosolic

denitrosylating activity was detected in numerous human cell types and rat tissues (Figure

23B). A subcellular fraction enriched with mitochondria also displayed denitrosylating

activity (Figure 22B).

Cytosolic denitrosylating activity was present in a fraction enriched for large

molecular size (> 10 kD), and was diminished after exposure to heat or trypsin (Figure

22C). Activity was also reduced after fractionation by size-exclusion chromatography

(Sephadex G-25), indicating a possible requirement for a small cofactor, and activity was

restored and potentiated by reduced nicotinamide adenine dinucleotide phosphate

(NADPH), but not nicotinamide dinucleotide (NADH), glutathione (GSH) or adenosine

triphosphate (ATP) (Figure 23C). Thus, the cytosolic denitrosylating activity exhibits

properties of an NADPH-dependent oxidoreductase.

We derived, by a four-step chromatographic purification from Jurkat cells, a

highly active fraction (designated fraction I) (Figure 23D and 24), whose activity was

dependent on a second fraction, added in limiting amounts (1:10) (designated fraction II)

Page 117: thesis mtf 1 - Duke University

102

(Figure 23D and 22D). Fraction I contained 8 proteins (Figure 22E), which were

identified by matrix-assisted laser desorption ionization time-of-flight mass spectrometry

(Figure 25). Of these, only thioredoxin-1 (Trx1) could be ascribed a redox-related

function. Recombinant thioredoxin reductase 1 (TrxR1) could substitute for fraction II,

fully reconstituting the denitrosylating activity of fraction I (Figure 22D).

Depletion of Trx1 from HeLa cells with short-interfering RNA (siRNA)

correlated with a loss of SNO-caspase-3 denitrosylating activity in vitro (Figure 26A and

27A), and denitrosylating activity was restored by adding back recombinant Trx1, but not

an active site mutant Trx (Trx1(C32S)) (Figure 27A). In contrast, siRNA-mediated

depletion of an additional member of the Trx family, thioredoxin-related protein 14

(TRP14) [173], had no effect on denitrosylating activity (Figure 27B). Similarly,

immunodepletion of Trx1 but not TRP14 abolished denitrosylating activity (Figure 27B).

A reconstituted Trx system (10 nM Trx and TrxR (Trx-TrxR) and including NADPH)

efficiently denitrosylated an excess of SNO-caspase-3 (Figure 26C). Denitrosylation by

Trx1 in the absence of TrxR1 was ineffective, but was restored when concentrations of

Trx1, but not Trx1(C32S), approached or exceeded that of SNO-caspase-3 (Figure 26D

and 27C), suggestive of single-turnover denitrosylation coupled to Trx1 oxidation.

Vicinal dithiols, as present in Trx (CXXC; where C represents Cys and X is

another amino acid), exhibit a denitrosylating activity in vitro [174], and various redox

enzymes, including dithiol proteins can catalyze denitrosylation reactions ex vivo [46,

175-177]. Trx may also break down SNOs within cells exposed to high concentrations of

a nitrosylating agent [178, 179], as might occur during a nitrosative stress. However, in

the context of signaling by endogenously-derived NO, Trx has been proposed to promote

Page 118: thesis mtf 1 - Duke University

103

S-nitrosylation of proteins (including caspase-3) [108, 180]. Thus in sum, it is not known

if Trx might have a role in mediating either basal or stimulus-induced denitrosylation in

vivo.

5.5 Trx1 mediates basal denitrosylation in intact cells

The extent of S-nitrosylation reflects the equilibrium between addition and

removal of NO groups. We employed RNA interference to assess if Trx governs basal S-

nitrosylation of caspase-3 in human B lymphocytes (10C9 cells), where most caspase-3 is

cytosolic and in the reduced form (i.e. only the minor population of caspase-3 associated

with the mitochondria is constitutively S-nitrosylated) [107]. Two non-overlapping

siRNAs efficiently decreased the amount of cellular Trx1, and amounts of endogenous

SNO-caspase-3 (assessed by the biotin-switch technique, in which S-nitrosylated

cysteines are selectively biotinylated [50, 116]) were increased in proportion to Trx1

depletion (Figure 28A). siRNA-mediated depletion of iNOS, or NOS inhibition with NG-

monomethyl-L-arginine (L-NMMA), prevented the increase in SNO-caspase-3 (Figure

29A and 29B). Depletion of TrxR1 in 10C9 cells (Figure 28A and 29C) and in human

embryonic kidney cells stably expressing a second NOS isoform, nNOS (HEK-nNOS)

(Figure 29D), also resulted in increased SNO-caspase-3, even more efficiently than did

depletion of Trx1. Thus, TrxR1 activity may be rate-limiting for denitrosylation of

cytosolic SNO-caspase-3 in vivo. Inhibition of Trx or TrxR might alter the cellular redox

balance (e.g. decrease intracellular GSH levels) and thus affect the abundance of SNO-

proteins independently of specific denitrosylating activity [116]. However, knockdown of

Page 119: thesis mtf 1 - Duke University

104

Trx1 or TrxR1 did not alter the levels of acid-soluble thiols (which consist primarily of

GSH) (Figure 29E).

Equivalent amounts of SNO-caspase-3 accumulated in HEK cells overexpressing

TrxR1 or empty vector shortly after treatment with S-nitrosocysteine (CysNO), a cell-

permeable S-nitrosylating compound (Figure 28B; 10 min). However, after 20 min, the

amount of SNO-caspase-3 was lower in TrxR1-transfected cells (Figure 28B). Therefore,

augmenting TrxR1 enhances SNO-caspase-3 denitrosylation. We examined the effects on

denitrosylation in intact cells of exposure to 1-chloro-2,4-dinitrobenzene (DNCB) or S-

triethylphosphine-gold(I)-2,3,4,6-tetra-O-acetyl-1-thio-β-D-glucopyranoside (auranofin),

two rapidly-acting and potent but structurally dissimilar inhibitors of TrxR [181, 182].

Treatment of 10C9 cells with either DNCB or auranofin increased the amount of SNO-

caspase-3 in cells exposed to CysNO (Figure 29F). Remarkably, in the absence of an

exogenous NO donor, cells treated with auranofin also accumulated SNO-caspase-3

(Figure 28C and 29G). We validated the identification of endogenous SNO in assays with

auranofin by showing the dependence of signals on ascorbate [50, 116] as well as by the

elimination of signals after UV irradiation, which liberates NO from SNO-proteins [116]

(Figure 28C and 29H), and we observed that auranofin did not alter the abundance of

acid-soluble thiols (Figure 29E). Thus, constitutive Trx1 activity apparently controls the

steady-state amount of caspase-3 S-nitrosylation, a role similar to that of phosphatases in

controlling basal phosphorylation.

To explore the role of Trx in reversing stimulus-coupled caspase-3 S-

nitrosylation, RAW264.7 macrophages were treated with lipopolysaccharide (LPS) plus

interferon-g (IFN-g), which induces iNOS. Whereas LPS/IFN-g or auranofin alone

Page 120: thesis mtf 1 - Duke University

105

weakly elicited S-nitrosylation of caspase-3, robust S-nitrosylation was seen when the

two treatments were combined (Figure 28D). Thus, S-nitrosylation of caspase-3 appears

to reflect dynamic turnover of the NO group, mediated by the opposing activities of NOS

and Trx-TrxR.

5.6 Denitrosylation proceeds, at least in part, via an intermolecular disulfide

The reduction of substrate disulfides by Trx involves a CXXC motif (Cys32 and

Cys35 in human Trx1) [183] (Figure 30A). A C35S mutant Trx can be used to trap

substrates in a mixed disulfide [184]. We developed a similar SNO-based trapping

strategy to examine the interaction of Trx with SNO-caspase-3. We transfected HEK

cells with caspase-3 and Trx1(C35S) fused to glutathione S-transferase (GST). After

treatment with CysNO, caspase-3 co-precipitated with Trx1(C35S) but not with wild-type

Trx1 (Figure 28E). The amount of caspase-3 trapped by Trx1(C35S) increased from 10 to

30 min after treatment with CysNO, during which time denitrosylation occurred (Figure

28F).

Additionally, in HEK cells transiently expressing iNOS, Trx(C35S) but not wild-

type Trx formed mixed disulfides with multiple proteins in an NO-dependent manner

(Figure 30D). Thus, Trx1 appears to interact with SNO-caspase-3 in vivo and to

denitrosylate via a mixed-disulfide intermediate (Figure 30B); a mechanism involving

transnitrosylation [178] is not, however, excluded (Figure 30C).

Page 121: thesis mtf 1 - Duke University

106

5.7 Trx1 is a broad-spectrum denitrosylase

Trx might serve as a denitrosylase for a broad spectrum of proteins. To identify

endogenously S-nitrosylated substrates of Trx under basal conditions, we inhibited Trx in

10C9 cells with auranofin, and identified caspase-9 and protein tyrosine phosphatase 1B

(PTP1B) as substrates for Trx-mediated denitrosylation (Figure 28G). The S-nitrosylated

form of glyceraldehyde-3-phosphate dehydrogenase (GAPDH), whose proapoptotic

function is induced by NO [27], was not detected in untreated cells, and did not

appreciably accumulate 2 hours after exposure to auranofin. However, treatment with

CysNO, representing a nitrosative stress, resulted in accumulation of SNO-GAPDH, and

the amounts of SNO-GAPDH were further increased markedly by auranofin (Figure

28H). Similarly, inhibition of TrxR or Trx increased the abundance of additional SNO-

proteins after exposure to CysNO (Figure 31A and 32B) as well as slowed their

denitrosylation (Figure 31C and 31D). Further, in HEK-nNOS cells, depletion of TrxR1

with siRNA increased the amounts of whole cell SNO (Figure 31E). Collectively, these

studies suggest that Trx-TrxR acts as a multi-substrate denitrosylase.

5.8 Trx2 mediates stimulus-induced denitrosylation of caspase-3 and apoptosis

We then asked whether the Trx system also mediates stimulus (Fas)-induced

denitrosylation, first described for mitochondria-associated SNO-procaspase-3 in 10C9

cells [25, 107]. Both siRNA-mediated depletion of TrxR2, the isoform of TrxR that is

localized selectively to mitochondria [185], and auranofin (which effectively inhibits

TrxR2 [186]), abrogated Fas-dependent denitrosylation of mitochondrial caspase-3

(Figure 32A, 32B and 33A). Thus, Fas-induced denitrosylation of mitochondrial SNO-

Page 122: thesis mtf 1 - Duke University

107

caspase-3 appears to be mediated by Trx2-TrxR2, a conclusion supported further by the

finding that Fas stimulation resulted in an increase in mitochondrial Trx2 (152% ± 12%

re. control, n=4) (Figure 33B). These results suggest the possibility that Trx2-mediated

denitrosylation (acting in concert with cleavage by initiator caspase(s)) may promote full

activation of caspase-3, and thereby facilitate apoptosis.

We further examined this possibility by assessing the effects of mitochondrial

TrxR2 knockdown or inhibition on two molecular events that characterize the execution

phase of Fas- and caspase-3- mediated apoptosis: caspase-3 activation and DNA

fragmentation [187]. In Fas-treated 10C9 cells, both depletion of TrxR2 with siRNA and

acute inhibition with auranofin reduced both the amount of cleaved, active caspase-3

(captured with biotin-Val-Ala-Asp(OMe) fluoromethyl ketone) (bVAD-FMK) (Figure

32C, 32D, 33C, and 34A) and caspase-3-like activity (cleavage of the tetrapeptide

DEVD) (Figure 33D). In support of these data, activation of caspase-8 (which lies

upstream of cytosolic caspase-3) was also diminished by TrxR2 inhibition (Figure 34A).

In contrast, depletion of TrxR1 had no appreciable effect on caspase activity (Figure 32D

and 33C). Furthermore, treatment of 10C9 cells with auranofin or knockdown of TrxR2

(but not TrxR1) with siRNA decreased DNA fragmentation by 45% ± 14 (n=4) (Figure

32E) and 34% ± 6 (Figure 32F and 34B), respectively. Although the precise sequence of

events subserving transmission of the NO-regulated apoptotic signal from mitochondrial

to cytosolic compartments remains to be elucidated fully (further discussed in Figure

34C), our findings suggest that denitrosylation of mitochondria-associated SNO-caspase-

3 by Trx2-TrxR2 promotes Fas-induced apoptosis. More generally, inhibition of Trx2-

TrxR2 provides a means to selectively manipulate mitochondria-associated caspase.

Page 123: thesis mtf 1 - Duke University

108

5.9 Discussion

Many SNO-proteins are present constitutively in low amounts, and both S-

nitrosylation and denitrosylation have been observed after stimulation of multiple classes

of receptors [21]. The present findings provide evidence for a specific enzymatic

mechanism of protein denitrosylation, acting in distinct cellular compartments to regulate

basal and stimulus-coupled protein denitrosylation, and complement the recent

description of S-nitrosylation/denitrosylation governed by GSNO reductase [46].

Discovery of mitochondria- and cytosol-specific denitrosylases not only elucidates the

mechanism by which Fas induces denitrosylation of mitochondria-associated caspases

selectively, but also indicates a function for Trx2 in amplifying apoptotic signals.

Previous studies have raised the idea that Trx1 may promote S-nitrosylation of

proteins through mechanisms involving non-active site thiols Cys69 or Cys73 [108, 180],

but these residues are not present in Trx2. Our finding that protein S-nitrosylation is

augmented by genetic or pharmacologic inhibition of either Trx1 or Trx2 suggests that a

major function of the Trx system (involving active-site Cys 32) is protein denitrosylation.

Most commonly, the Trx system is characterized as playing a highly-conserved role in

protection from cellular oxidative stress. The present findings indicate that SNO-protein

denitrosylation (and thus regulation of NO-based signaling) represents an additional

function of the Trx system in metazoan organisms, in which several classes of SNO-

regulating enzymes—NO synthases, GSNO reductases and Trx-TrxRs—are expressed

ubiquitously across cell types.

Page 124: thesis mtf 1 - Duke University

109

Figure 22. Identification of Trx1 as a caspase-3 denitrosylase. (A) Strategy for biochemical identification of SNO-caspase-3 denitrosylase. Hexahistidine-tagged recombinant procaspase-3 was S-nitrosylated at the active site Cys and thereby inhibited before immobilization on nickel-coated microwell plates. Denitrosylation by cellular extracts was revealed by restoration of protease activity using N-acetyl-(Asp-Glu-Val-Asp)-7-amido-4-methylcoumarin (Z-DEVD-AMC) as substrate. (B) Caspase-3 activity after 30 min incubation of SNO-caspase-3 with mitochondria-enriched or cytosolic fractions (100 µ g protein) prepared from Jurkat cells. (C) Caspase-3 activity after incubation of SNO-caspase-3 with cytosolic fraction prepared from Jurkat cells (Cyto) or with cytosolic extract boiled for 15 min, treated with trypsin (2 µg; 30 min), or enriched for high MW species by ultrafiltration (10 kDa cutoff) or for low MW species (the flow through after ultrafiltration). (D) Caspase-3 activity after incubation of SNO-caspase-3 with fractions obtained during the purification procedure (see Materials and Methods for details), and/or with recombinant rat TrxR1 (5 nM). (E) Representative SDS-PAGE gel corresponding to the chromatographic purification scheme of SNO-caspase-3 denitrosylating activity (see figure 24 and 25). Data (B, C and D) are presented as mean ± SEM; n = 3

Page 125: thesis mtf 1 - Duke University

110

Figure 23. Characterization of an enzymatic activity that denitrosylates SNO-caspase-3. (A) Reactivation of SNO-caspase-3 protease activity by cell cytosol. Immobilized SNO-caspase-3 (~100 nM) was incubated for 30 min with a cytosolic extract prepared from Jurkat cells. At left, caspase activity was determined using Z-DEVD-AMC. At right, SNO content was assayed using Hg-coupled photolysis-chemiluminescence. a.u., arbitrary unit. (B) Caspase activity after 30 min incubation of SNO-caspase-3 with cytosolic fractions (100 µg protein) prepared from human cells or rat tissues. HAEC, primary human aortic endothelial cells. (C) SNO-caspase-3 was incubated with the cytosolic fraction from Jurkat cells, or with cytosolic fractions following size-exclusion chromatography through Sephadex G-25 (high MW fraction) and supplemented with GSH (0.5 mM), ATP (10 mM), NADH (10 mM) or NADPH (10 mM). (D) The procedure used to partially purify a SNO-caspase-3 denitrosylase. Data in A-C are presented as mean ± SEM; n ≥ 3.

Page 126: thesis mtf 1 - Duke University

111

Figure 24. Partial purification of SNO-caspase-3 denitrosylating activity.

Page 127: thesis mtf 1 - Duke University

112

Figure 25. Proteins in “Fraction I” identified by MALDI-TOF MS.

Page 128: thesis mtf 1 - Duke University

113

Figure 26. Modulation of Trx1 activity determines caspase-3 denitrosylation. (A) Caspase-3 activity was determined (with Z-DEVD-AMC) after 30 min incubation of SNO-caspase-3 (~100 nM) with a cytosolic fraction prepared from untreated HeLa cells or from cells that were transfected for 3 days with siRNA for Trx1. (B) Caspase-3 activity was determined after 30 min incubation of SNO-caspase-3 with HeLa cytosolic extract or cytosol that had been depleted of Trx1 or TRP14 using specific anti-Trx1 or anti-TRP14 antibodies. (C) Caspase-3 activity was determined after 30 min incubation of SNO-caspase-3 (~100 nM) with recombinant human Trx1 (10 nM) and/or recombinant rat TrxR1 (10 nM), and with NADPH (100 µM). (D) Caspase-3 activity after 30 min incubation with recombinant E. coli Trx1. Data in (A-D) are presented as mean ± SEM; n = 3.

Page 129: thesis mtf 1 - Duke University

114

Figure 27. The active site of Trx1 is critical for denitrosylation. (A) Caspase-3 activity was determined after incubation of SNO-caspase-3 with cytosolic fractions prepared from HeLa cells that were transfected with siRNA for Trx1 (oligo 2) or with control siRNA, or with Trx1-depleted cytosol supplemented with 100 nM recombinant wild-type Trx1 (Trx-WT) or with Trx1(C32S). (B) Caspase-3 activity after incubation of SNO-caspase-3 with a cytosolic fraction prepared from HeLa cells that were transfected with siRNA for TRP14 or with control siRNA. (C) Caspase-3 activity after incubation of SNO-caspase-3 with 100 nM recombinant Trx-WT or with Trx(C32S).

Page 130: thesis mtf 1 - Duke University

115

Figure 28. Trx1-TrxR1 mediates protein denitrosylation in vivo. (A) 10C9 cells were transfected for 3 days with siRNA specific for Trx1 or TrxR1, or with control RNA. SNO-caspase-3 was assayed by biotin-switch. The histogram summarizes results (mean ± SEM) of three experiments. (B) HEK cells were transfected for 24 hours with TrxR1 before treatment with CysNO (200 µM). Whole cell extracts were prepared 10 or 20 min after CysNO exposure and SNO-caspase-3 was assayed by biotin-switch. (C) 10C9 cells were treated with auranofin (2 µM) for the times indicated and SNO-caspase-3 was assayed by biotin-switch. To verify biotin labeling specificity, the effect of omitting ascorbate in the assay is shown. (D) RAW264.7 cells were treated with auranofin (1 mM, 1 hour) or with LPS (1 mg/ml)/IFNg (10 ng/ml) for 16 hours, or LPS/IFNg plus auranofin (added for the last hour), and SNO-caspase-3 was assayed by biotin-switch. (E) SNO-dependent interaction of Trx1 and caspase-3. HEK cells were co-transfected with caspase-3 and either GST-tagged wild-type Trx1, GST-tagged Trx1(C35S) or empty GST vector before treatment with CysNO (500 µM; 10 min). Pull-down from lysates was with GSH-agarose. (F) HEK cells were treated as in E and lysed at different times after CysNO exposure. Pull-down of proteins was as in E (top). SNO-caspase-3 was assessed by biotin-switch (bottom). (G, H) 10C9 cells were treated with auranofin (2 µM) or vehicle (DMSO) for 2 hours and S-nitrosylation of endogenous caspase-9, PTP1B and GAPDH was assessed by biotin-switch. For GAPDH (H), CysNO (500 mM; 20 min) was added following auranofin treatment. Results in A-H are representative of 3 experiments.

Page 131: thesis mtf 1 - Duke University

116

Figure 29. Trx-TrxR inhibition increases the amount of SNO-caspase-3 in vivo. (A) 10C9 cells were transfected for 3 days with siRNA for Trx1 and siRNA for iNOS, as indicated. SNO-caspase-3 was assayed in whole cell extracts by biotin switch. (B) 10C9 cells were transfected with siRNA for Trx1 or with control siRNA, followed by treatment with L-NMMA (5 mM, 2 hours). SNO-caspase-3 was assayed in whole cell extracts by biotin switch. (C) 10C9 cells were transfected for 3 days with siRNA for TrxR1 (oligo 2) or with control siRNA, and the amount of SNO-caspase-3 was determined by biotin switch. (D) Neuronal NOS-expressing HEK cells (HEK-nNOS) were transfected for 3 days with siRNA for TrxR1 or with control siRNA, and for one day with caspase-3 cDNA. The amount of SNO-caspase-3 was determined by biotin switch. (E) 10C9 or HEK-nNOS cells were transfected with siRNA for Trx1 or TrxR1 or treated for 2 hours with auranofin. The amount of acid soluble thiols present in cell extracts was evaluated by DTNB colorimetric assay. (F) 10C9 cells were treated with DNCB (30 µM) or auranofin (2 µM) for 1 hour followed by exposure to CysNO (500 µM; 30 min). SNO-caspase-3 was assayed by biotin switch. Control samples were treated with decayed (old) CysNO. (G) HEK-nNOS cells were transfected for 1 day with caspase-3 cDNA, followed by 1 hour treatment with auranofin. The amount of SNO-caspase-3 was determined by biotin switch. (H) 10C9 cells were treated with auraonofin (2 µM; 2 hours). Lysates were then exposed or not to UV irradiation for 3 min and SNO-caspase-3 was assayed by biotin switch.

Page 132: thesis mtf 1 - Duke University

117

Figure 30. Interaction of Trx with SNO-proteins. (A) Reaction mechanism of disulfide reduction by Trx. In the initial step, the N-terminal Cys thiolate (within the CXXC motif of Trx) attacks the target disulfide leading to formation of a mixed disulfide intermediate, which is followed by release of substrate as the disulfide is transferred to Trx via attack by the C-terminal Cys. (B) A possible reaction scheme for SNO-caspase denitrosylation by Trx-TrxR. (C) An alternative reaction scheme for SNO-caspase denitrosylation by Trx-TrxR that is based on transnitrosylation. (D) HEK cells were co-transfected (24 hours) with pIRES-iNOS and either GST-tagged-Trx1-WT or GST-tagged Trx1(C35S). Pull-down from lysates was with GSH-agarose. The complexes were resolved by electrophoresis under non-reducing conditions, and detected by western blotting with an anti-GST specific antibody.

Page 133: thesis mtf 1 - Duke University

118

Figure 31. Trx-TrxR inhibition increases the amount of SNO-proteins in vivo. (A) 10C9 cells were treated with auranofin (2 µM; 1 hour) followed by exposure to CysNO (500 µM; 30 min). Assessment of SNO in whole cell extracts was done by DAF-2 assay. CysNO rapidly decayed and thus resulted in insignificant SNO readings in this assay. (B) 10C9 cells were treated as in A. Protein S-nitrosylation in whole cell extracts was assessed by biotin-switch. (C) HEK cells were treated with DNCB (30 µM; 1 hour) then exposed to CysNO (500 µM). Levels of whole cell SNO-protein were determined using the DAF-2 assay at different intervals following CysNO exposure. Data are presented as mean ± SEM; n = 3. (D) HEK cells were transfected with siRNA for Trx1 or with control siRNA. After 3 days cells were treated with CysNO and analyzed for SNO content as in C. Data are presented as mean ± SEM; n = 3. (E) Wild-type HEK cells (HEK-WT) or HEK-nNOS cells were transfected for 3 days with siRNA for TrxR1 or with control siRNA. Assessment of SNO in whole cell extracts was done by DAF-2 assay. Data are presented as mean ± SEM; n=3. *P < 0.05 by ANOVA.

Page 134: thesis mtf 1 - Duke University

119

Figure 32. The mitochondrial thioredoxin system mediates Fas-induced denitrosylation of mitochondria-associated SNO-caspase-3, and promotes apoptotic signaling. (A) 10C9 cells were transfected for 3 days with siRNA for TrxR2 or with control RNA before exposure to anti-Fas CH11 antibody (50 ng/ml) for 2 hours. The amount of SNO-caspase-3 in a subcellular fraction enriched for mitochondria was evaluated by biotin-switch. Results are mean ± SEM of 4 experiments. *P < 0.05 by ANOVA. (B) 10C9 cells were left untreated or treated with auranofin for 1 hour, followed by Fas receptor stimulation and evaluation of mitochondrial SNO-caspase-3 as in A. Results are the mean ± SEM of 3 experiments. *P < 0.05 by ANOVA. (C) 10C9 cells treated as in A were lysed in the presence of bVAD-FMK (10 µM), an affinity ligand for active caspase. Cleaved caspase-3 (< 20 kDa) was assessed by immunoblotting with anti-caspase-3 antibodies in lysates and following purification of active, bVAD-FMK-bound caspases with streptavidin-agarose (pull-down). Results are representative of 3 experiments. (D) 10C9 cells were transfected with siRNA specific for TrxR1, TrxR2 or with control RNA. Cells were left untreated or treated with auranofin (1 µM; 1 hour) followed by treatment with anti-Fas CH11 antibody (100 ng/ml; 2 hours). Caspase-3 cleavage and activation were assessed as in C. Results (Fas-activated samples) are mean ± SEM of 3 experiments. *P < 0.05, **P < 0.01 by ANOVA. (E, F) 10C9 cells were treated with auranofin, or transfected with siRNA specific for TrxR1 or TrxR2 as in A-D. DNA fragmentation was assessed 6 hours after treatment with anti-Fas CH11 antibody. Results (Fas-activated samples) are mean ± SEM of 4 experiments. *P < 0.05 by ANOVA.

Page 135: thesis mtf 1 - Duke University

120

Figure 33. Effects of TrxR2 inhibition on Fas-induced denitrosylation and activation of caspase-3. (A) 10C9 cells were transfected for 3 days with siRNA for TrxR2 (oligo 2) or with control siRNA before exposure to anti-Fas CH11 antibody (50 ng/ml) for 2 hours. The amount of SNO-caspase-3 in a subcellular fraction enriched for mitochondria was evaluated by biotin-switch. (B) 10C9 cells were treated with Fas ligand (50 ng/ml) for 2 hours, and the amounts of Trx2 and of the mitochondrial protein dihydrolipoamide dehydrogenase (DLDH) in a fraction enriched for mitochondria were determined by immunoblotting. The histogram summarizes results of 4 experiments (mean ± SEM); * p<0.05. (C) 10C9 cells were transfected with siRNA for TrxR1 (oligo 2), TrxR2 (oligo 2) or with control siRNA. Cells were treated with anti-Fas CH11 antibody (100 ng/ml; 2 hours) and caspase-3 cleavage was assessed by immunoblotting. (D) 10C9 cells were treated with auranofin (1 µM, 1 hour) followed by exposure to anti-Fas CH11 antibody for 2 hours. Caspase-3-like/DEVDase activity in whole cell lysates was measured by a fluorogenic assay using DEVD-AMC as a substrate. Data are presented as mean ± SEM; n = 3.

Page 136: thesis mtf 1 - Duke University

121

Figure 34. Effects of TrxR2 inhibition on Fas-induced activation of caspase-3 and caspase-8 and on DNA fragmentation. (A) 10C9 cells were transfected for 3 days with siRNA for TrxR2 or with control siRNA before exposure to anti-Fas CH11 antibody for 2 hours. Cleavage of caspase-3 and of caspase-8 was assessed in whole cell lysates by immunoblotting. (B) 10C9 cells transfected with siRNA for TrxR1 (oligo 2) or TrxR2 (oligo 2) or with control siRNA. DNA fragmentation was assessed 6 hours after Fas treatment. Results (Fas-activated samples) are mean ± SEM of 3 experiments. *P < 0.05 by ANOVA with Tukey post hoc test. Within-gel cropping of images is indicated by a white line. (C) A proposed model for involvement of caspase-3 denitrosylation in Fas-signaling. Engagement of Fas receptor triggers cytoplasmic and mitochondrial events that

combine to induce apoptosis. Previous findings have demonstrated that activation of caspase-8 results in cleavage and activation of cytosolic caspase-3, inducing DNA fragmentation and cell death. In some cells, this cytosolic pathway only weakly activates caspase-3 and apoptosis is dependent upon a mitochondrial amplification loop, in which cleavage of proapoptotic Bid, followed by its translocation to the mitochondria, triggers the release of cytochrome c and activation of caspase-9. Our findings point to an additional mechanism whereby mitochondria may amplify the apoptotic signal, in which Fas up-regulates mitochondrial Trx2-TrxR2 (step 1) thereby inducing denitrosylation of mitochondria-associated SNO-procaspase-3 (step 2), required for activation. Initiator caspase-dependent cleavage of denitrosylated caspase-3 (in or at mitochondria or following translocation to cytosol) completes the process of activation (step 3). The newly activated caspase-3 may then cleave cytosolic and/or mitochondrial substrates and may promote the activation of initiator caspases (steps 4 and 5), amplifying the apoptotic cascade. Dashed arrows indicate indirect or incompletely elucidated interactions or sequences of events.

Page 137: thesis mtf 1 - Duke University

122

Chapter 6. Thioredoxin Interacting Protein (Txnip) is a Feedback Regulator of

S-Nitrosylation

This chapter is based on a manuscript that is currently being prepared for submission to

the Proceedings of the National Academy of Sciences.

Forrester, M.T., Seth, D., Hausladen, A., Eyler, C.E., Matsumoto, A., Benhar, M., Marshall, H.E., & Stamler, J.S. Thioredoxin interacting protein (Txnip) is a feedback regulator of S-nitrosylation Proc Nat Acad Sci USA, In Preparation, (2009).

Page 138: thesis mtf 1 - Duke University

123

6.1 Summary

Nitric oxide exerts a plethora of biological effects via protein S-nitrosylation, a

redox-based reaction that converts a protein Cys thiol to an S-nitrosothiol. Despite the

wealth of efforts demonstrating a role for nitric oxide synthases in driving S-nitrosylation,

relatively few studies have identified the cellular components that mediate or regulate

protein S-denitrosylation. Recently, the thioredoxin system—a ubiquitous Cys-based

oxidoreductase—was shown to mediate both basal and stimulus-coupled protein S-

denitrosylation. Here we demonstrate that an endogenous thioredoxin inhibitor,

thioredoxin-interacting protein (Txnip), is downregulated by endogenously synthesized

nitric oxide. This, in turn, affords thioredoxin an optimal environment for protein S-

denitrosylation and allows the cell to survive nitrosative stresses that are otherwise

detrimental. These findings demonstrate that enzymatic pathways of denitrosylation are

dynamically regulated systems and reveal new potential strategies for pharmacological

manipulation of S-nitrosylation.

Page 139: thesis mtf 1 - Duke University

124

6.2 Introduction

It has become increasingly appreciated that protein S-nitrosylation—the covalent

attachment of a nitroso group to a cysteine thiol side chain—is a principle mechanism by

which nitric oxide (NO) modulates numerous cellular functions and phenotypes [21,

188]. These include G-protein-coupled receptor signaling [110, 189, 190], death receptor-

mediated apoptosis [24, 25, 153, 167, 191, 192], glutamate-dependent neurotransmission

[84, 193-195], vesicular trafficking [55, 196-198], stimulation of prostaglandin synthesis

[85, 199, 200], and the unfolded protein response [52]. In addition, aberrant S-

nitrosylation is implicated in pathologies such as tumor initiation and growth [9, 201-

204], neurodegeneration [51, 52, 205-207] and malignant hyperthermia [208].

The three isoforms of nitric oxide synthase (i.e. nNOS/NOS1, iNOS/NOS2,

eNOS/NOS3) are well-established stimuli for S-nitrosylation, and numerous studies have

demonstrated that their localization and/or NO-producing activities are critical for S-

nitrosylation of target proteins. For example, binding of iNOS to COX2 is critical for S-

nitrosylation and activation of the latter [85], while the subcellular localization of eNOS

is another established determinant of protein S-nitrosylation [74, 209].

Despite the plethora of efforts to elucidate the components that promote S-

nitrosylation via NO production or localization, the intracellular factors that mediate

denitrosylation have only recently gained attention as potential regulators of overall S-

nitrosylation. By analogy to phosphorylation (where kinases and phosphatases together

regulate phosphorylation), the steady-state level of S-nitrosylation is the net difference

between nitrosylation and denitrosylation. The activities of denitrosylation systems may

therefore also play a role in regulating the net levels of protein S-nitrosylation.

Page 140: thesis mtf 1 - Duke University

125

One of the first described pathways for enzymatic denitrosylation is S-

nitrosoglutathione (GSNO) reductase (GSNOR) [45, 46]. This NADH-dependent

oxidoreductase specifically reduces GSNO, as protein-SNOs are not direct substrates

(Figure 3). Since GSH/GSNO and protein-SH/protein-SNOs are in a constant equilibrium

via rapid transnitrosation (nitroso transfer) reactions [210, 211], the activity of GSNOR

does in fact indirectly lower protein S-nitrosylation via reduction of GSNO [47, 212].

6.3 Thioredoxin is an Enzymatic Denitrosylase

Recent studies have demonstrated that thioredoxin (Trx) is a broad-spectrum

intracellular denitrosylase [153, 178, 179], a finding that has generated a surge of interest

regarding the overall role of Trx in regulating protein-SNOs and their established

signaling pathways. The Trx system was originally described as a disulfide reduction

pathway; the mechanism is illustrated in Figure 35. The mechanism of denitrosylation,

however, is less well understood. Figure 36 illustrates the two likely mechanisms of

denitrosylation by the Trx1 system, which differ by whether nitroso transfer chemistry or

disulfide formation predominate. In our published study, cytosolic thioredoxin

(thioredoxin-1; Trx1) was shown to mediate constitutive basal denitrosylation, thus

maintaining low levels of S-nitrosylation in various cell types (Figure 37) [153]. The

mitochondrial isoform of thioredoxin (thioredoxin-2; Trx2) was also shown to mediate

stimulus-coupled denitrosylation of caspase-3 [153], an established system in which

denitrosylation is coupled to activation of the pro-apoptotic death receptor family (i.e.

Fas, TNF receptor), leading to efficient induction of programmed cell death [25, 167].

Thioredoxin has therefore been shown to mediate both basal and stimulus-coupled

Page 141: thesis mtf 1 - Duke University

126

denitrosylation, though Trx was originally described as a protein disulfide reductase and

electron donor for ribonucleotide reductase [213]. The overall thioredoxin system is

composed of two proteins: 1) the 12 kD Trx itself, which employs an active site dithiol

motif to reduce target proteins, and 2) the 50 kD flavoprotein thioredoxin reductase

(TrxR), which employs an active site selenocysteine and NADPH to reduce oxidized

Trx1 and thus complete the catalytic cycle. For denitrosylation, both components and

their requisite cofactors appear critical [153, 178].

6.4 Nitric Oxide Synthesis and Nitrosative Stress

A separate though intriguing issue is how mammalian cells can produce NO and

impose a nitrosative stress on pathogens without themselves succumbing to the same

stress. In numerous microbes, the NO-inducible flavohemoglobin has been shown to

consume NO [36, 37, 39] and contribute to nitrosative stress resistance [40, 44, 214],

which strengthens the notion that NO is employed by hosts as a part of their antimicrobial

armamentarium. Since the mammalian host is the primary source of NO, one might

expect that an equivalent mammalian NO-consuming activity would be

counterproductive to the antimicrobial functions of NO (and therefore not exist).

However, an inducible denitrosylation pathway could serve as an anti-nitrosative

protection system, and we therefore hypothesized that cells may augment their Trx

system to achieve this goal. Presented here are data suggesting that the Trx system is

directly activated by NO-dependent repression of thioredoxin-interacting protein, an

established inhibitor of thioredoxin.

Page 142: thesis mtf 1 - Duke University

127

6.5 Thioredoxin Interacting Protein: An Introduction

Thioredoxin-interacting protein (Txnip, VDUP1, TBP-2) was originally described

as a vitamin-D3-upregulated protein, isolated from an HL60 cell-derived cDNA library

[215]. Subsequent studies by Yodoi and coworkers identified an interaction between

Txnip and Trx1 via a yeast two-hybrid assay using Trx1 as bait [216], an interaction

which leads to the inhibition of Trx1 in vivo [216-219]. Junn et al. demonstrated that

Txnip-mediated inhibition of Trx1 potentiates damage from reactive oxygen species

(ROS) [217], and a wealth of recent studies have shown that Txnip is a negative regulator

of insulin function, thus promoting diabetic phenotypes [218, 220-223]. Rather

independently, Schulze et al. showed that exogenous GSNO attenuates the expression of

Txnip [224], though these authors did not examine the effects of endogenous NO

production on Txnip or elucidate a physiological function for GSNO-dependent Txnip

suppression. Nonetheless, the report from Schulze et al. was both inciting and

informative for our studies presented here, which examine the interplay between

endogenous NO synthesis, Txnip, protein denitrosylation and cellular resistance to

nitrosative stress.

6.6 Materials and Methods

Materials and Reagents – All materials were from Sigma unless otherwise

indicated. 3H-thymidine was from Cambridge Isotopes. 1400W, DETA-NO and ODQ

were from Cayman Chemical. Sources of antibodies were: monoclonal α-Txnip/Vdup1

(MBL International, Catalog # K0205-3); monoclonal α-iNOS (BD Biosciences, Catalog

Page 143: thesis mtf 1 - Duke University

128

# 610431); monoclonal α-GAPDH (Millipore, Catalog # MAB374); human-specific

monoclonal α-Trx1 (BD Biosciences, Catalog # 559969); monoclonal anti-GST (Santa

Cruz, Catalog # sc-138); mouse-specific polyclonal α-Trx1 (Cell Signaling, Catalog #

2298); polyclonal α-TrxR1 (Abcam, Catalog # ab16840-100); polyclonal α-Prx1

(Upstate, Catalog # 07-609); polyclonal α-Prx2 (Lab Frontier, Catalog # LF-PA0007).

All antibodies were used in 5% non-fat dry milk in TBST. CysNO was synthesized as

described [116] and used immediately.

Mammalian Cell Culture and Transfection – All cells were cultured at 37 °C in a

5% CO2 atmosphere. Where indicated, the culture incubator (Sanyo MCO-18M) was

purged with moistened N2 gas to achieve lower O2 concentrations. Cell lines were

obtained via the Duke Cell Culture Facility and grown in DMEM with 10% FBS, 100

U/ml penicillin and 100 µg/ml streptomycin. Cells were transfected with Lipofectamine

2000 (Invitrogen) per the manufacturer’s instructions. To generate stable cell

transfectants, HEK293 cells in a 6 well dish were transfected with 3 µg of the indicated

pCDNA3.1 plasmid and 7.5 µl Lipofectamine 2000, and the media was changed after 8 h.

Twenty four hours later, G418 (Invitrogen) was added to a final concentration of 1 mg/ml

of active antibiotic. After 2 weeks, individual colonies were picked by gentle aspiration

with a wide boar pipette tip and transferred to 24 well dishes. Cells were maintained with

G418 for another 3 weeks. Bone marrow derived macrophages (BMDMs) were isolated

as described [225], and grown in DMEM containing 20% L-929 conditioned media, 10%

FBS, 100 U/ml penicillin and 100 µg/ml streptomycin.

Page 144: thesis mtf 1 - Duke University

129

Cloning and DNA Manipulation – All PCRs were performed with Advantage Taq

DNA polymerase (Clontech) and products verified by DNA sequencing (Duke DNA

Sequencing Facility). The cDNA encoding human iNOS (Genbank L09210) was excised

from pCWori-iNOS (a kind gift from Dennis Stuehr, Cleveland Clinic) by restriction

digestion at the NdeI and XbaI sites. This 3.5 kb fragment was ligated into pDNR-1r

donor plasmid (Clontech). The iNOS E377Q mutation was introduced via site-directed

mutagenesis with the Quickchange kit (Stratagene) according to the manufacturer’s

instructions using the following primers: 5’-GTACATGGGCACACAGATCGGAGTCC-

3’ and 5’-GGACTCCGATCTGTGTGCCCATGTAC-3’. The human iNOS WT and

E377Q cDNAs were then transferred into pLP-IRES-neo via the Creator system

(Clontech) according to the manufacturer’s instructions. An IMAGE clone of mouse

Txnip cDNA in pCMV-SPORT6 (BC031850) was acquired from Open Biosystems. To

generate stable cell transfectants overexpressing Txnip, the cDNA of Txnip was

subcloned into the 5’-EcoRI and 3’-XhoI sites in pCDNA3.1 via PCR with the following

primers: 5'-

TATTGAATTCACCATGGACTACAAAGACGATGACGACAAGGTGATGTTCAAG

AAGATCAAGTC-3' and 5'-

TAATCTCGAGTCATTTCCTGCAGGCTCACTGCACGTT-3'. Mammalian expression

plasmids encoding GST-tagged Trx1 (pEBG-Trx1) were kindly provided by Ki-Sun

Kwon (Korea Research Institute of Bioscience and Biotechnology, Daejeon, South

Korea) [226]. Txnip reporter plasmids were created by PCR amplification of the Txnip

promoter from human genomic DNA (isolated from HEK293 cells) and ligation into

pGL4.14 at the 5’-XhoI and 3’-HindIII sites. To generate the 845 bp fragment (524 bp of

Page 145: thesis mtf 1 - Duke University

130

promoter and 321 bp of 5’-UTR), the following primers were used: 5'-

ATTACTCGAGCAGCCCCAAACCTGAAAGTATTCTTGGAGC-3' and 5'-

ATATAAGCTTCTGAGTTGGTTTTAAGAGTTAGAAATGACGG-3'. To generate

only the 524 bp of promoter (lacking the 5’-UTR), the following primers were used: 5'-

ATTACTCGAGCAGCCCCAAACCTGAAAGTATTCTTGGAGC-3' and 5'-

AATTAAGCTTAAGCACTCCTTTGGAGAA-3'.

Quantitative RT-PCR of Txnip – Following the treatment of cells as indicated,

total RNA was extracted with the RNeasy kit (Qiagen) and 500 ng of product was

converted to cDNA with random hexamer oligonucleotides and MuMLV reverse

transcriptase (Invitrogen) according to the manufacturer’s instructios. Gene specific

primers (described below) were used for RT-PCR in a MyiQ RT-PCR Detection System

(BioRad) using 1 µl of cDNA as template and 2X iQ SYBR green supermix (BioRad).

The expression of β-actin RNA in each sample was used to normalize the expression of

Txnip. The size of each PCR product was verified by agarose gel electrophoresis. Fold

change in expression was calculated using the comparative Ct method. For murine

(RAW264.7) cells, the following primers were used for Txnip: 5’-

CATGAGGCCTGGAAACAAAT-3’ and 5’-ACTGGTGCCATTAGGTCAGG-3’. For

murine β-actin, the following primers were used: 5’-AGCCATGTACGTAGCCATCC-3’

and 5’-CTCTCAGCTGTGGTGGTGAA-3’. For human (HEK293) cells, the following

primers were used for Txnip: 5’-TTCGGGTTCAGAAGATCAGG-3’ and 5’-

TTGGATCCAGGAACGCTAAC-3’. For human β-actin, the following primers were

Page 146: thesis mtf 1 - Duke University

131

used: 5'-TCAAGATCATTGCTCCTCCTGAGC-3' and 5'-

TTGCTGATCCACATCTGCTGGAAG-3'.

Detection of Protein-SNOs with the Biotin Switch Technique (BST) – The BST

was performed as described [116], with minor modifications as described below. For

determination of total protein-SNOs, the streptavidin-agarose was eluted under non-

reducing conditions by boiling in 60 µl of 25 mM HEPES, 0.1 mM EDTA, 0.01 mM

neocuproine, pH 8.0, containing 1% SDS. Eluants were resolved by SDS-PAGE,

transferred to a nitrocellulose membrane and incubated overnight with streptavidin-HRP

(Amersham/GE) at 1:1000 dilution in 5% non-fat dry milk in TBST.

Measurement of Cellular Damage and Proliferation – Cellular release of LDH

was measured with the Cytotox-96 assay kit (Promega) as described by the

manufacturer’s instructions, except quantification was performed by normalization to a

standard curve of purified LDH (Sigma). 3H-thymidine (Perkin Elmer) uptake was

measured by diluting the 3H-thymidine stock solution (1 µCi/µl) 1:250 into culture

media. After 4 h, cells were gently collected with 1 ml PBS, centrifuged at 5000 g, and

the remaining PBS was removed. The cell pellet was washed once with 10%

trichloroacetic acid (TCA) and then fixed for 1 h at 4 °C in 10% TCA. This material was

centrifuged at 5000 g, supernatant was discarded and the pellet was resuspended in 1.0 ml

of 0.2 M NaOH, incubated overnight and measured by scintillation counting.

Page 147: thesis mtf 1 - Duke University

132

6.7 Txnip Expression is Negatively Regulated by Endogenous Nitric Oxide

Given our hypothesis that cells manipulate the Trx system during conditions of

endogenous nitrosative stress, we initially sought to examine the expression of various

redox-active proteins in resting vs. NO-producing cells. We employed the RAW264.7

murine macrophage cell line, which readily expresses iNOS following exposure to LPS

and interferon-y (IFN-γ) [227, 228] and is an established model system for endogenous

nitrosative stress [27, 117, 229]. As shown in Figure 38, stimulation of macrophages with

LPS and IFN-γ led to increased iNOS expression, though the protein levels of the

cytosolic Trx1 or TrxR1 were not significantly changed. In contrast, the expression of

Txnip was significantly decreased following cytokine stimulation and NO-production,

which was partially reversed by co-incubation with the iNOS-specific inhibitor 1400W.

As assessed by nitrite accumulation in the media, iNOS activity was approximately 90%

inhibited by 1400W, suggesting that the residual iNOS activity may be responsible for

the continued Txnip suppression under these conditions. Also observed was increased

expression of peroxiredoxin-1 (Prx1), a Trx-dependent H2O2 reductase (peroxidase),

following cytokine stimulation. This induction was unaffected by iNOS inhibition with

1400W, suggesting that Prx1 expression is not driven by NO, though Diet et al. observed

iNOS-dependent Prx1 induction in primary murine macrophages [230]. One possible

explanation for this discrepancy is that 1400W does not fully inhibit iNOS (~10%

activity remains) and this remaining activity may be sufficient to drive Prx1 induction,

which is abrogated when examining primary macrophages from iNOS-/- mice.

To assess whether NO-mediated Txnip suppression is restricted to murine cells,

we examined Txnip expression in human embryonic kidney (HEK293) cells transfected

Page 148: thesis mtf 1 - Duke University

133

with cDNA encoding human iNOS. As demonstrated in Figure 39, overexpression of

iNOS attenuated the expression of endogenous Txnip, but had no significant effect on

several other redox active proteins. This effect was reversed by 1400W and an inactive

iNOS mutant (E377Q) was also without effect on Txnip expression. Importantly, the

E377Q mutation of iNOS abrogates the ability of iNOS to bind L-Arg (substrate), though

it has no effects on other iNOS activities such as cytochrome c reduction or NADPH

oxidation [231]. These results demonstrate that 1) the effect of iNOS on Txnip

suppression is in fact due to NO production and not simply the presence of iNOS protein

or other potential iNOS-mediated side reactions (e.g. O2- production) and 2) the

machinery by which NO acts to suppress Txnip expression exist both in murine and

human cells, and thus is not likely to be a species-specific phenomenon.

To explore the temporal properties of Txnip suppression, we performed a time

course with cytokine-stimulated macrophages. As shown in Figure 40, RAW264.7

macrophages exhibit Txnip suppression as early as 3 h post-stimulation. Inhibition of

iNOS with 1400W, however, increases Txnip expression at later time points (> 9 h),

during which time iNOS expression is prevalent. Therefore, Txnip undergoes two phases

of suppression: an early, iNOS-independent phase and a later, iNOS-dependent phase.

Given our interest in the Trx-mediated denitrosylation (and the role of Txnip in this

process), we have focused our efforts on the iNOS-dependent time points. Further, the

iNOS-mediated effect seems necessary to maintain continued repression of Txnip during

the later phases (> 9 h) of macrophage stimulation (Figure 40). To further investigate the

physiological relevance of iNOS-dependent Txnip suppression, we examined the levels

of Txnip expression in bone-marrow derived macrophages (BMDMs) from wild-type and

Page 149: thesis mtf 1 - Duke University

134

iNOS-/- mice. Following bone marrow harvesting and culture for 7 days in the presence of

M-CSF, BMDMs were either untreated or stimulated with LPS and IFN-γ (500 ng/ml and

100 U/ml, respectively) for 12 h, lysed and analyzed by SDS-PAGE and western blotting.

As shown in Figure 41, BMDMs from wild-type mice exhibited cytokine-dependent

expression of iNOS and pronounced repression of Txnip, though BMDMs from iNOS-/-

mice exhibited only a slight decrease in Txnip expression. These experiments collectively

demonstrate that endogenous NO production, either from cytokine-stimulated iNOS

expression or transfection of iNOS cDNA, directly suppresses the expression of Txnip in

both human and murine systems.

6.8 Txnip Suppression is Replicated by Exogenous Nitric Oxide

To examine the effects of exogenous NO donor on Txnip expression, HEK293

cells were treated for 18 h with various concentrations of DETA-NO, a slow-releasing

NO donor (t1/2 ~ 20 h). As shown in Figure 42, the effects of iNOS transfection or

cytokine stimulation on Txnip expression were closely replicated by DETA-NO

treatment, suggesting that NO itself is sufficient for driving Txnip suppression. In this

experiment, HEK293 cells were employed in lieu of RAW264.7 macrophages because

the latter exhibited a small degree of iNOS induction following treatment with DETA-

NO, which we felt was a complicating factor when attempting to examine the effects of

an exogenous NO donor. Collectively, these experiments establish that both endogenous

and exogenously generated NO lowers the expression of Txnip protein.

Page 150: thesis mtf 1 - Duke University

135

6.9 Txnip Suppression is Independent of O2 and cGMP

Though most mammalian cell culture is performed under atmospheric O2 (21%),

this high level of O2 is infrequently physiological, as most perivascular tissues experience

2-5% O2 [232, 233]. Since high O2 may impose an oxidative stress, affect signaling via

the hypoxia-inducible factor (HIF) pathway [234] or alter NO-based redox reactions

[235-237], we sought to examine the effects of these somewhat artificially hyperoxic

conditions on NO-mediated Txnip suppression. As shown in Figure 43, the degree of

iNOS-mediated Txnip suppression is unaffected by a more normoxic environment (i.e.

2% O2), suggesting that NO inhibits Txnip expression under normal physiological O2

concentrations and is not simply promoted by inadvertent effects of O2 tension on NO

biology.

An established pathway of NO-based signaling involves the activation of soluble

guanylyl cyclase (sGC), which converts GTP to cGMP [238, 239]. To determine whether

NO-mediated suppresion of Txnip is driven by sGC and the production of cGMP, we

examined the effects of the competitive sGC inhibitor, ODQ [240], on Txnip expression.

As demonstrated by Figure 44A, stimulation-dependent Txnip suppression was

unaffected ODQ, suggesting that NO inhibits the expression of Txnip irrespective of sGC

activity and cGMP production. In a complimentary experiment, the effects of 8-bromo-

cGMP (a cell permeable and stable analogue of cGMP [241]) on Txnip expression was

investigated in RAW264.7 macrophages. As shown in Figure 44B, 8-bromo-cGMP does

not affect Txnip expression, though cytokine stimulation over the same time period (16 h)

leads to a profound suppression of Txnip. Collectively, these experiments demonstrate

that endogenously generated NO lowers the expression of Txnip protein via a cGMP- and

Page 151: thesis mtf 1 - Duke University

136

O2-independent pathway, arguing for a redox-based mechanism and function for Txnip

suppression.

6.10 Txnip Suppression Involves mRNA Downregulation

The level of Txnip protein could theoretically be modulated by NO at many

points in the pathway of expression. These include transcriptionally (i.e. promoter

activity), post-transcriptionally (e.g. mRNA stability or localization), translationally, or

post-translationally (i.e. protein stability). To examine the effects of NO on Txnip mRNA

levels, RAW264.7 macrophages were either untreated or stimulated in the presence or

absence of 1400W, RNA was isolated and subjected to quantitative RT-PCR (qPCR). As

shown by Figure 45A, cytokine stimulation leads to a robust decrease in Txnip mRNA to

approximately 10% of the basal levels. At time points when iNOS expression is prevalent

(e.g. 9 h), Txnip mRNA suppression is partially reversed by 1400W. These mRNA levels

correlate with Txnip protein expression at these time points (Figure 38), suggesting that

the effects of NO on Txnip protein expression occur, at least in part, at the level of Txnip

mRNA abundance. To perform a similar experiment in a human system, we examined the

effects of iNOS transfection on Txnip mRNA in HEK293 cells. As shown in Figure 45B,

active wild-type iNOS led to a reduction in Txnip mRNA, though the inactive iNOS

mutant (E377Q) was without effect.

6.11 Txnip Protein Undergoes Rapid Intracellular Turnover

Despite these effects of iNOS activity on Txnip mRNA, overall protein levels of

Txnip should only be sensitive to decreases in mRNA levels if Txnip protein normally

Page 152: thesis mtf 1 - Duke University

137

undergoes rapid turnover (or else Txnip protein would remain unchanged despite a

decrease in cognate mRNA). To examine the stability of Txnip protein, RAW264.7 cells

were incubated with cycloheximide (to inhibit ribosomal protein synthesis) and Txnip

protein was assessed by western blot. As shown in Figure 46, Txnip protein is highly

unstable and undergoes complete turnover within 90 min, though other proteins such as

GAPDH are highly stable during this interval. This short half-life of Txnip protein may

be due to either ubiquitin-dependent or independent proteolysis. Collectively, these

findings suggest that Txnip protein is sufficiently unstable such that mRNA fluctuations

may readily affect Txnip protein levels.

6.12 Nitric Oxide Inhibits the Txnip Promoter

As Txnip mRNA is without any identifiable regulatory regions in the 3’-UTR

[242], we hypothesized that the effects of NO on Txnip mRNA abundance would likely

occur at the level of promoter activity or possibly, through modulation of the 5’-UTR. To

examine this hypothesis more fully, we constructed firefly luciferase reporter plasmids

containing 524 bp of the Txnip promoter with or without the 321 bp of 5’-UTR (845 bp

of DNA total). HEK293 cells were transfected with these reporter plasmids, as well as

empty control plasmid or plasmid encoding human iNOS (WT and E377Q mutants).

Firefly luciferase activity was normalized with an internal control Renilla luciferase

plasmid to prevent any effects of cell viability or transfection efficiency on Txnip

reporter measurements. As shown in Figure 47, the Txnip reporter activity was inhibited

by both exogenous NO donor and transfection of active iNOS, though the iNOS(E377Q)

mutant was without effect. DETA-NO was employed as an NO donor because its half life

Page 153: thesis mtf 1 - Duke University

138

is approximately 20 h at 37 °C, and it therefore generates a sustained NO release rate

with low overall NO concentrations (i.e. nM steady-state NO). Importantly, the NO-

dependent suppressive effects were observed both with and without the 5’-UTR of the

Txnip gene. These data demonstrate that NO represses Txnip expression at the level of

promoter activity, though other effects such as altered mRNA stability or localization

have certainly not been excluded. Importantly, Schulze et al. also observed suppression

of Txnip promoter activity following treatment of cells with exogenous GSNO [224],

which is overall similar to our observations with exogenous NO donor (DETA-NO).

6.13 Txnip Repression Protects from Nitrosative Stress

To examine whether NO-dependent Txnip repression serves to protect the cell

from nitrosative stress (e.g. by relieving inhibition of Trx), we examined the effects of

Txnip overexpression on NO-mediated cell damage. HEK293 cells were transfected with

combinations of plasmids encoding Txnip and/or iNOS (total transfected DNA was kept

constant with empty plasmids). As shown in Figure 48, Txnip overexpression alone

caused no detectable release of lactate dehydrogenase (LDH), an intracellular enzyme

and marker of cell damage, into the culture media. In contrast, iNOS overexpression

alone induced readily detectable LDH release, suggesting the iNOS itself can damage

cells. The combination of Txnip and iNOS overexpression, however, synergistically

drove LDH release, suggesting that iNOS toxicity is highly potentiated by Txnip. Further,

these data suggest that endogenous Txnip suppression by NO may serve a cytoprotective

function. To determine whether Txnip overexpression potentiated iNOS-dependent

damage via activation of iNOS activity, media nitrite (NO2-, an end product of NO

Page 154: thesis mtf 1 - Duke University

139

synthesis) was measured. As shown in Figure 49, NO2- levels were not increased, but

rather slightly decreased by Txnip co-expression, likely due to decreased cell viability

during co-transfection of Txnip and iNOS. Importantly, these data demonstrate that the

effects of Txnip are independent of iNOS activity, and thus downstream of NO

production. Though LDH release is a reliable marker of cell damage, it does not reveal

information regarding cellular growth. To determine the effects of Txnip and iNOS on

cellular growth and DNA synthesis, 3H-thymidine uptake was assayed in HEK293 cells

overexpressing Txnip and/or iNOS. As shown in Figure 50, cells transfected with Txnip

alone exhibited a slight decrease in 3H-thymidine uptake, consistent with previous reports

that Txnip attenuates 3H-thymidine uptake [217, 218]. The effects of iNOS on 3H-

thymidine uptake however, were again highly potentiated by Txnip co-transfection,

suggesting that Txnip modulates both NO-dependent cytostasis and cellular damage

(LDH release). To directly demonstrate the effects of iNOS and Txnip on cellular

morphology and viability, cells from the LDH experiment were visualized by light

microscopy. As shown in Figure 51, transfection of iNOS and Txnip together causes

morphological changes consistent with cell damage, which is much less pronounced with

iNOS transfection alone. Cells transfected with both iNOS and Txnip exhibit decreased

intercellular adhesion and increased swelling consistent with gross cellular damage.

6.14 Txnip Binds to the Active Site Cys32 of Trx1

Previous studies have employed site-directed mutagenesis to demonstrate that

Txnip binds to the active site of Trx1, which consists of the nucleophilic Cys32 and

“resolving” Cys35. These studies, however, compared either WT Trx1 against a double

Page 155: thesis mtf 1 - Duke University

140

mutant (C32S/C35S) [217] or employed a C73S mutant for all pulldown experiments

[243]. As the Trx1 Cys73 residue contains a surface-exposed and reactive thiol [244, 245],

we sought to confirm that the Txnip-Trx1 interaction is in fact via the active site

nucleophilic thiol (Cys32) of Trx1 (and not the allosteric/non-catalytic Cys73), which

would explain the inhibitory properties of Txnip on Trx1 activity. To this end, HEK293

cells were transfected with cDNAs encoding GST-tagged Trx1 (GST-Trx1) and

examined for interactions with endogenous Txnip by affinity isolation with GSH-agarose

and western blotting. Importantly, all isolated proteins were eluted only with reductant

(not denaturant/detergent), thus confirming that protein-protein interactions are via a

disulfide linkage. As shown in Figure 52, the active site Cys32 was necessary for the

Trx1-Txnip interaction, as the C32S mutant was unable to bind Txnip. These data also

explain why the Trx1-Txnip interaction is inhibitory for Trx1, as Txnip directly binds to

the active site Cys32 of Trx1 and prevents reduction or denitrosylation of target substrates.

6.15 Txnip Potentiates Intracellular Protein S-Nitrosylation

Given the recent demonstrations that Trx is a constitutively active denitrosylase

[153, 178, 179], we sought to examine the effects of Txnip on intracellular S-

nitrosylation. As shown in Figure 53A, HEK293 cells transfected with iNOS exhibit

increased intracellular levels of protein-SNOs, as measured by Hg-coupled photolysis-

chemiluminescence, which employs UV-based photolysis of the S-NO bond and ozone-

based detection of the release NO [75]. Co-transfection of Txnip with iNOS, however,

further increases protein S-nitrosylation by approximately 40% (n = 6, p < 0.05). As

shown in Figure 53B, analogous results were obtained with the biotin switch technique

Page 156: thesis mtf 1 - Duke University

141

(BST), in which protein-SNOs are chemically converted to disulfide-linked biotin

moieties and detected with avidin-based reagents [50, 116]. Numerous protein-SNOs are

increased by Txnip co-expression, suggesting that Txnip is a broad-spectrum regulator of

protein S-nitrosylation. Since S-nitrosylated GAPDH (GAPDH-SNO) is an established

mediator of NO-based cell death [27, 246], we reasoned that Txnip may regulate cellular

tolerance to nitrosative stress via modulation of GAPDH-SNO. As shown in Figure 54, S-

nitrosylation of GAPDH in iNOS-transfected HEK293 cells was highly potentiated by

Txnip co-transfection. These findings suggest that Txnip is a broad-spectrum regulator of

S-nitrosylation, mediating its effects, at least in part, via regulation of GAPDH-SNO.

6.16 Discussion

Protein S-nitrosylation and denitrosylation are highly dynamic events, and the

molecular mechanisms governing both processes are currently being elucidated. For

protein denitrosylation, the major enzymatic pathways are S-nitrosoglutathione reductase

(GSNOR) and the Trx system. Both of these redox systems have been shown to drive

denitrosylation under basal conditions, and it has been demonstrated that Trx mediates

caspase-3 denitrosylation in response to death-receptor stimulation (e.g. Fas, TNF-α

receptor). However, whether either of these systems are themselves regulated,

particularly in response to NO production, has not yet been demonstrated. Here we

describe a novel NO-regulated feedback system to control S-nitrosylation, which involves

the NO-dependent transcriptional repression of Txnip (illustrated in Figure 55). This

pathway is independent of cGMP and operates under physiological O2 concentrations.

Importantly, these observations demonstrate that Trx1-dependent denitrosylation is a

Page 157: thesis mtf 1 - Duke University

142

highly regulated system, and that repression of a Trx1-inhibitor (Txnip) serves to

facilitate denitrosylation and protect from nitrosative stress.

Despite our demonstration that Trx-mediated denitrosylation is itself highly

regulated, numerous fundamental questions remain to be pursued in future studies. For

example, which transcription factor(s) senses NO and drives Txnip suppression? As sGC

appears uninvolved in this event, it is possible that the molecular mechanism by which

Txnip is suppressed involves a redox-signaling mechanism (e.g. S-nitrosylation of a

transcription factor or other trans-acting component).

Of particular interest is how Txnip can form a stable disulfide with Trx1 within

the highly reducing cytosolic environment. As the redox potential of mammalian

cytoplasm is ~ -300 mV [247, 248], intra and intermolecular disulfides are considered

thermodynamically unstable. Therefore, the Txnip-Trx1 intermolecular disulfide is likely

to be thermodynamically forbidden, implying that its stability is supported by a kinetic

barrier to reduction. This kinetic barrier could be provided by steric factors that make the

Txnip-Trx1 disulfide bond inaccessible to cellular reductants (e.g. GSH). Future studies

focused on the structural aspects of Txnip (and its association with Trx1) will hopefully

reveal how the Txnip-Trx1 disulfide bond can survive in a highly reducing environment.

The precise mechanism by which nitrosative stress drives cell death, regulated in

part via Txnip suppression, is also an active area of interest. Our studies demonstrate that

Txnip is a broad-spectrum regulator of S-nitrosylation, as one might expect considering

its target (Trx1) is a broad-spectrum denitrosylase [153]. It does indeed protect GAPDH

from S-nitrosylation, which has been shown to activate the acetyltransferase p300 and

consequently induce p53 and PUMA [249]. Despite these demonstrated effects of Txnip

Page 158: thesis mtf 1 - Duke University

143

on iNOS-dependent S-nitrosylation and nitrosative stress, Txnip may also modulate the

plethora of eNOS- and nNOS-dependent signaling pathways. Ongoing efforts in our lab

are aimed at this hypothesis, and it is our hope that the current report serves as a proof-of-

principle that Trx-mediated denitrosylation is dynamically regulated via Txnip

expression.

Page 159: thesis mtf 1 - Duke University

144

Figure 35. The reaction pathway of disulfide reduction by the thioredoxin system. In the initial step, the N-terminal Cys32 thiolate (within the CXXC motif of Trx) attacks the target disulfide leading to formation of a mixed disulfide intermediate, which is followed by release of substrate as the disulfide is transferred to Trx via attack by the C-terminal “resolving” Cys35.

Page 160: thesis mtf 1 - Duke University

145

Figure 36. Two potential reaction pathways of protein denitrosylation by the thioredoxin system. (Left) A plausible reaction scheme for caspase denitrosylation by Trx-TrxR based on nucleophilic attack of the Cys32 thiolate on the sulfur atom of the caspase-SNO with concomitant release of nitroxyl anion (NO-). Following formation of an intermolecular disulfide, Cys35 resolves the complex into reduced caspase substrate and oxidized Trx1. (Right) An alternative reaction scheme for SNO-caspase denitrosylation by Trx-TrxR based on transnitrosation. In this mechanism, the nitroso group (NO+) is transferred between the caspase and Cys32 of Trx1, followed by release of nitroxyl and formation of an intramolecular disulfide on Trx1. In both mechanisms, the oxidized Trx1 is re-reduced by thioredoxin reductase (TrxR).

Page 161: thesis mtf 1 - Duke University

146

Figure 37. The thioredoxin system mediates protein denitrosylation in intact cells. (A) Human B-lymphocytes (10C9 cells) were transfected for 3 days with siRNA specific for Trx1 or TrxR1 or with control siRNA. Caspase-3-SNO was assayed by the biotin switch technique (BST). The histogram summarizes results (mean ± SEM) of three experiments. (B) HEK293 cells were transfected for 24 hours with either empty plasmid or plasmid encoding TrxR1 prior to treatment with CysNO (200 µM). Whole-cell extracts were prepared 10 or 20 min after CysNO exposure, and caspase-3-SNO was assayed by the BST. (C) B-lymphocytes (10C9 cells) were treated with the TrxR inhibitor auranofin (2 µM) for the times indicated, and caspase-3-SNO was assayed by biotin switch. To verify biotin labeling specificity for SNO, the effect of omitting ascorbate in the assay is shown.

Page 162: thesis mtf 1 - Duke University

147

Figure 38. Thioredoxin-interacting protein (Txnip) is suppressed in murine macrophages in response to endogenous NO production. Macrophages (RAW264.7 cells) were left untreated or exposed to LPS and IFN-γ (500 ng/ml and 100 U/ml, respectively) for the indicated time intervals. Where indicated, 100 µM 1400W (iNOS inhibitor) was added at the time of stimulation, resulting in approximately 90% inhibition of iNOS activity (as assessed by NO2

- accumulation in the culture media). Cellular extracts were subjected to SDS-PAGE and western blotting for the indicated proteins, revealing a pronounced repression of Txnip upon macrophage stimulation, which is partly reversed by inhibition of iNOS.

Page 163: thesis mtf 1 - Duke University

148

Figure 39. Thioredoxin-interacting protein (Txnip) is suppressed in HEK293 cells transfected with iNOS in an NO-dependent manner. Human embryonic kidney (HEK293) cells were transfected with either empty pLP, pLP-iNOS(WT) or pLP-iNOS(E377Q). After 20 h, cells were harvested and expression levels of various redox proteins were assessed by western blotting. As indicated, transfection of active, wild-type iNOS led to decreased expression of Txnip, which was reversed by iNOS inhibition and not observed upon transfection of an inactive iNOS mutant (E377Q).

Page 164: thesis mtf 1 - Duke University

149

Figure 40. A time-course study in stimulated murine macrophages reveals both early and late phases of Txnip suppression, which are NO-independent and NO-dependent, respectively. RAW264.7 macrophages were stimulated with LPS and IFN-γ for the indicated time periods in the absence or presence of 1400W, and Txnip/iNOS expression were assessed by SDS-PAGE and western blotting. As shown, Txnip suppression is evident as early as 3 h post-stimulation, though iNOS expression is barely detectable at this time point. Further, 1400W has no effect on this phenomenon. On the contrary, by 9 h post-stimulation, iNOS expression is prevalant and Txnip suppression is partially reversed by 1400W, suggesting that this phase is indeed dependent on NO production.

Page 165: thesis mtf 1 - Duke University

150

Figure 41. Primary macrophages from iNOS-/- mice fail to suppress Txnip in response to cytokine stimulation. Bone marrow derived macrophages (BMDMs) from wild-type and iNOS-/- mice were left untreated or stimulated with LPS and IFN-γ (100 ng/ml and 100 U/ml, respectively) for 12 h and the expression of Txnip/iNOS were assessed by SDS-PAGE and western blotting. In wild-type BMDMs, iNOS is highly induced following stimulation and Txnip undergoes suppression. In BMDMs derived from iNOS-/- mice, however, this effect is largely attenuated.

Page 166: thesis mtf 1 - Duke University

151

Figure 42. Suppression of Txnip is replicated by exogenously supplied NO. HEK293 cells were incubated with various concentrations of DETA-NO, a slow-releasing NO-donor (t1/2 ~ 20 h) and Txnip expression was assessed by SDS-PAGE and western blotting. This experiment demonstrates that exogenous NO mimics the effects of cytokine stimulation and endogenous iNOS activity on Txnip, thus independently confirming the role of NO in Txnip suppression.

Page 167: thesis mtf 1 - Duke University

152

Figure 43. NO-dependent Txnip suppression is independent of ambient O2 concentrations. RAW264.7 macrophages were stimulated with LPS/IFN-γ for 16 h in either typical experiment O2 (i.e. 21%), or a more physiological environment of 2% O2, and the expression of Txnip/iNOS were assessed by SDS-PAGE and western blotting. As indicated, the concentration of O2 was without effect, suggesting that the effects of NO on Txnip expression are not secondary to the relatively hyperoxic typical experimental conditions (i.e. 21% O2).

Page 168: thesis mtf 1 - Duke University

153

Figure 44. NO-dependent Txnip suppression is independent of soluble guanylyl cyclase and cGMP. (A) RAW264.7 macrophages were stimulated with LPS/IFN-γ for 16 h in the absence or presence of ODQ, a potent and irreversible inhibitor of soluble guanylyl cyclase, and the expression of Txnip/iNOS were assessed by SDS-PAGE and western blotting. (B) RAW264.7 macrophages were stimulated with LPS/IFN-γ for 16 h or treated with various concentrations of 8-bromo-cGMP, a cell-permeable analogue of cGMP and direct activator of protein kinase G. As indicated, neither treatment affected the results of cytokine stimulation and NO production, suggesting that NO does not mediate the suppression of Txnip via soluble guanylyl cylase and cGMP production. These results suggest a redox-dependent mechanism of and function for Txnip suppression (e.g. S-nitrosylation).

Page 169: thesis mtf 1 - Duke University

154

Figure 45. NO-dependent Txnip suppression involves a decrease in Txnip mRNA in both murine and human cells. (A) RAW264.7 macrophages were stimulated with LPS/IFN-γ for the indicated time periods, total cellular RNA was extracted and subjected to quantitative RT-PCR (qPCR). Txnip RNA levels were normalized to β-actin. Where indicated, iNOS was inhibited by ~90% with 100 µM 1400W (** p < 0.01). (B) HEK293 cells were transfected with either empty pLP-IRES-Neo, pLP-iNOS(WT) or pLP-iNOS(E377Q) for 20 h, and cellular RNA was subjected to qPCR. Txnip RNA levels were normalized to β-actin. These data suggest that iNOS mediates Txnip suppression, at least in part, via lowering the levels of Txnip mRNA.

Page 170: thesis mtf 1 - Duke University

155

Figure 46. A time-course of Txnip protein stability in RAW264.7 macrophages reveals this protein is highly unstable. Intracellular protein synthesis was inhibited with the ribosomal inhibitor cycloheximide (CHX; 100 µM). Cells were harvested at the indicated time points and Txnip protein levels were assessed by SDS-PAGE and western blotting. As indicated, Txnip protein is highly unstable relative to other proteins (e.g. GAPDH) and undergoes complete turnover within 90 min. The high instability of Txnip likely allows the protein levels to drop following a decrease in cognate mRNA. If the protein were highly stable, it would be difficult to understand how a drop in mRNA might lead to rapid decreases in protein abundance.

Page 171: thesis mtf 1 - Duke University

156

Figure 47. NO-dependent Txnip suppression involves transcriptional deactivation of the Txnip promoter as assessed by a luciferase reporter assay. HEK293 cells were transfected with pGL4.14-Txnip (524 bp of promoter -/+ 321 bp of 5’-UTR) along with pRL-SV40 for internal normalization of luciferase activity. Cells were either treated with DETA-NO for 12 h or co-transfected with pLP-iNOS(WT) or pLP-iNOS(E377Q). In each sample, total transfected DNA was equalized with empty pLP-IRES-Neo plasmid. Twenty four hours after transfection, cells were harvested and assayed for firefly and Renilla luciferase activities. These results demonstrate that the human Txnip promoter is inhibited by active iNOS or exogenous NO, suggesting that endogenously produced NO lowers Txnip protein and mRNA levels via shutting down the Txnip promoter.

Page 172: thesis mtf 1 - Duke University

157

Figure 48. Txnip facilitates nitrosative stress in iNOS-transfected HEK293 cells, as assessed by LDH release into the media. HEK293 cells in 6 well dishes were transfected with either empty pCMV (0, 2 or 4 µg), pCMV-Txnip (0, 2 or 4 µg), empty pLP (3 µg) or pLP-iNOS(WT) (3 µg). Eight hours after transfection, media was changed. After another 28 h, media was collected and lactate dehydrogenase (LDH) activity in the media was measured along with LDH standards.

Page 173: thesis mtf 1 - Duke University

158

Figure 49. Txnip-dependent nitrosative stress is not secondary to increased iNOS activity. HEK293 cells in 6 well dishes were transfected with either empty pCMV (0, 2 or 4 µg), pCMV-Txnip (0, 2 or 4 µg), empty pLP (3 µg) or pLP-iNOS(WT) (3 µg). Eight hours after transfection, media was changed. After another 28 h, media was collected and nitrite (NO2

-) in the media was measured by Griess assay along with NO2- standards.

Page 174: thesis mtf 1 - Duke University

159

Figure 50. Txnip facilitates nitrosative stress in iNOS-transfected HEK293 cells, as assessed by cellular 3H-thymidine uptake. HEK293 cells in 6 well dishes were transfected with either empty pCMV (0, 2 or 4 µg), pCMV-Txnip (0, 2 or 4 µg), empty pLP (3 µg) or pLP-iNOS(WT) (3 µg). After 20 h, cellular uptake of 3H-thymidine was assayed as a measurement of cellular growth. In each experiment, total DNA and transfection reagent were kept constant between samples. DPM, disintegrations per minute.

Page 175: thesis mtf 1 - Duke University

160

Figure 51. Cellular morphology during Txnip-driven nitrosative stress. HEK293 cells from which LDH release was assayed (see Figure 48) were visualized by light microscopy. Shown are cells transfected with pLP-iNOS(WT) and either 4 µg of empty pCMV or pCMV-Txnip plasmid. As indicated, cells that lack Txnip exhibit normal morphology when transfected with iNOS. However, cells transfected with both Txnip and iNOS exhibit grossly visible changes typical of cellular damage (e.g. decreased cell-cell contact, cell blebbing and swelling, decreased adhesion).

Page 176: thesis mtf 1 - Duke University

161

Figure 52. Thioredoxin-interacting protein (Txnip) binds to the active site Cys32 of Trx1 via a disulfide bond. HEK293 cells were transfected with pEBG-GST-Trx1(WT) and (C32S) active site mutant for 20 h, lysed and Trx1 was immunoprecipitated with GSH-agarose. Beads were eluted with 50 mM DTT to release all disulfide-linked proteins, and bound Txnip was assessed by SDS-PAGE and western blotting. As indicated, the active site Cys32 of Trx1 was required for disulfide formation with Txnip in intact cells. These data demonstrate how Txnip inhibits Trx1 (i.e. via directly binding to the active site Cys32).

Page 177: thesis mtf 1 - Duke University

162

Figure 53. Txnip facilitates general protein S-nitrosylation as assessed by Hg-coupled photolysis-chemiluminescence (“Nitrolite”) and the biotin switch technique (BST). (A) HEK293 cells stably transfected with either pCDNA3.1 or pCDNA3.1-Txnip were co-transfected with pLP or pLP-iNOS. After 20 h, cells were harvested and proteins isolated by size-exclusion chromatography with P-6 desalting gel. Protein-SNO was measured by Hg-coupled photolysis-chemiluminescence, which employs UV-photolysis of the S-NO bond and detection of NO by chemiluminescent reaction with ozone. GSNO was employed for SNO standards. Protein content was measured by BCA assay (** p < 0.01). (B) Protein-SNOs were assayed by the biotin switch technique (BST) following co-transfections as in (A). Total protein biotinylation (i.e., SNO) were detected by immunodetection with avidin-HRP. Indicated is a single representative experiment where several intermediate samples (transfected with less iNOS plasmid) have been removed.

Page 178: thesis mtf 1 - Duke University

163

Figure 54. Txnip potentiates the S-nitrosylation of GAPDH, a major cellular effector of NO-dependent cell death. HEK293 cells were transfected as in the previous figure and the S-nitrosylation of GAPDH was detected by the BST with western blotting for GAPDH. As indicated, GAPDH-SNO is readily detected following co-transfection of iNOS and Txnip, though neither transfection alone has an appreciable affect on the S-nitrosylation of GAPDH.

Page 179: thesis mtf 1 - Duke University

164

Figure 55. A scheme of Txnip function in protein S-nitrosylation and denitrosylation. Protein S-nitrosylation is frequently promoted by NOS activity (e.g., iNOS in a cytokine-stimulated macrophage). Protein-SNOs are constitutively denitrosylated by thioredoxin-1 (Trx1), via either nitroso-transfer or intermolecular disulfide formation. Both mechanistic routes lead to oxidized Trx1, which is in turn reduced by thioredoxin reductase (TrxR1). Independent of these events, NO production negatively regulates the expression of Txnip, an established inhibitor of Trx1. By relieving the inhibition of Trx1, cells are able to tolerate elevated production of NO without experiencing excessive protein S-nitrosylation and nitrosative stress, though the role of this pathway in SNO-based signal transduction pathways is by no means excluded.

Page 180: thesis mtf 1 - Duke University

165

References

1. Fridovich, I., Oxygen toxicity: a radical explanation. J Exp Biol, 1998. 201(Pt 8): p. 1203-9.

2. Imlay, J.A., Cellular defenses against superoxide and hydrogen peroxide. Annu Rev Biochem, 2008. 77: p. 755-76.

3. Stuehr, D.J., et al., Update on mechanism and catalytic regulation in the NO synthases. J Biol Chem, 2004. 279(35): p. 36167-70.

4. Babior, B.M., NADPH oxidase. Curr Opin Immunol, 2004. 16(1): p. 42-7.

5. Bedard, K. and K.H. Krause, The NOX family of ROS-generating NADPH oxidases: physiology and pathophysiology. Physiol Rev, 2007. 87(1): p. 245-313.

6. Nathan, C., Nitric oxide as a secretory product of mammalian cells. FASEB J, 1992. 6(12): p. 3051-64.

7. Nathan, C. and M.U. Shiloh, Reactive oxygen and nitrogen intermediates in the relationship between mammalian hosts and microbial pathogens. Proc Natl Acad Sci U S A, 2000. 97(16): p. 8841-8.

8. Cals-Grierson, M.M. and A.D. Ormerod, Nitric oxide function in the skin. Nitric Oxide, 2004. 10(4): p. 179-93.

9. Li, F., et al., Regulation of HIF-1alpha stability through S-nitrosylation. Mol Cell, 2007. 26(1): p. 63-74.

10. Rai, R.M., et al., Impaired liver regeneration in inducible nitric oxide synthasedeficient mice. Proc Natl Acad Sci U S A, 1998. 95(23): p. 13829-34.

11. Sinz, E.H., et al., Inducible nitric oxide synthase is an endogenous neuroprotectant after traumatic brain injury in rats and mice. J Clin Invest, 1999. 104(5): p. 647-56.

Page 181: thesis mtf 1 - Duke University

166

12. Cramer, J.P., et al., iNOS promoter variants and severe malaria in Ghanaian children. Trop Med Int Health, 2004. 9(10): p. 1074-80.

13. Lopansri, B.K., et al., Low plasma arginine concentrations in children with cerebral malaria and decreased nitric oxide production. Lancet, 2003. 361(9358): p. 676-8.

14. Ohashi, J., et al., Significant association of longer forms of CCTTT Microsatellite repeat in the inducible nitric oxide synthase promoter with severe malaria in Thailand. J Infect Dis, 2002. 186(4): p. 578-81.

15. Parikh, S., G. Dorsey, and P.J. Rosenthal, Host polymorphisms and the incidence of malaria in Ugandan children. Am J Trop Med Hyg, 2004. 71(6): p. 750-3.

16. Imlay, J.A., S.M. Chin, and S. Linn, Toxic DNA damage by hydrogen peroxide through the Fenton reaction in vivo and in vitro. Science, 1988. 240(4852): p. 640-2.

17. Park, S., X. You, and J.A. Imlay, Substantial DNA damage from submicromolar intracellular hydrogen peroxide detected in Hpx- mutants of Escherichia coli. Proc Natl Acad Sci U S A, 2005. 102(26): p. 9317-22.

18. Schapiro, J.M., S.J. Libby, and F.C. Fang, Inhibition of bacterial DNA replication by zinc mobilization during nitrosative stress. Proc Natl Acad Sci U S A, 2003. 100(14): p. 8496-501.

19. Woodmansee, A.N. and J.A. Imlay, A mechanism by which nitric oxide accelerates the rate of oxidative DNA damage in Escherichia coli. Mol Microbiol, 2003. 49(1): p. 11-22.

20. Cary, S.P., et al., Nitric oxide signaling: no longer simply on or off. Trends Biochem Sci, 2006. 31(4): p. 231-9.

21. Hess, D.T., et al., Protein S-nitrosylation: purview and parameters. Nat Rev Mol Cell Biol, 2005. 6(2): p. 150-66.

Page 182: thesis mtf 1 - Duke University

167

22. Russwurm, M. and D. Koesling, Guanylyl cyclase: NO hits its target. Biochem Soc Symp, 2004(71): p. 51-63.

23. Zheng, M. and G. Storz, Redox sensing by prokaryotic transcription factors. Biochem Pharmacol, 2000. 59(1): p. 1-6.

24. Kim, J.E. and S.R. Tannenbaum, S-Nitrosation regulates the activation of endogenous procaspase-9 in HT-29 human colon carcinoma cells. J Biol Chem, 2004. 279(11): p. 9758-64.

25. Mannick, J.B., et al., Fas-induced caspase denitrosylation. Science, 1999. 284(5414): p. 651-4.

26. Mitchell, D.A., S.U. Morton, and M.A. Marletta, Design and characterization of an active site selective caspase-3 transnitrosating agent. ACS Chem Biol, 2006. 1(10): p. 659-65.

27. Hara, M.R., et al., S-nitrosylated GAPDH initiates apoptotic cell death by nuclear translocation following Siah1 binding. Nat Cell Biol, 2005. 7(7): p. 665-74.

28. Hara, M.R., M.B. Cascio, and A. Sawa, GAPDH as a sensor of NO stress. Biochim Biophys Acta, 2006. 1762(5): p. 502-9.

29. Zhu, L., C. Gunn, and J.S. Beckman, Bactericidal activity of peroxynitrite. Arch Biochem Biophys, 1992. 298(2): p. 452-7.

30. Koppenol, W.H., et al., Peroxynitrite, a cloaked oxidant formed by nitric oxide and superoxide. Chem Res Toxicol, 1992. 5(6): p. 834-42.

31. Pacher, P., J.S. Beckman, and L. Liaudet, Nitric oxide and peroxynitrite in health and disease. Physiol Rev, 2007. 87(1): p. 315-424.

32. Fridovich, I., Superoxide radical and superoxide dismutases. Annu Rev Biochem, 1995. 64: p. 97-112.

33. Fridovich, I., Superoxide anion radical (O2-.), superoxide dismutases, and related matters. J Biol Chem, 1997. 272(30): p. 18515-7.

Page 183: thesis mtf 1 - Duke University

168

34. Imlay, J.A., How oxygen damages microbes: oxygen tolerance and obligate anaerobiosis. Adv Microb Physiol, 2002. 46: p. 111-53.

35. Imlay, J.A., Pathways of oxidative damage. Annu Rev Microbiol, 2003. 57: p. 395-418.

36. Crawford, M.J. and D.E. Goldberg, Role for the Salmonella flavohemoglobin in protection from nitric oxide. J Biol Chem, 1998. 273(20): p. 12543-7.

37. Gardner, P.R., et al., Nitric oxide dioxygenase: an enzymic function for flavohemoglobin. Proc Natl Acad Sci U S A, 1998. 95(18): p. 10378-83.

38. Hausladen, A., A. Gow, and J.S. Stamler, Flavohemoglobin denitrosylase catalyzes the reaction of a nitroxyl equivalent with molecular oxygen. Proc Natl Acad Sci U S A, 2001. 98(18): p. 10108-12.

39. Hausladen, A., A.J. Gow, and J.S. Stamler, Nitrosative stress: metabolic pathway involving the flavohemoglobin. Proc Natl Acad Sci U S A, 1998. 95(24): p. 14100-5.

40. Liu, L., et al., Protection from nitrosative stress by yeast flavohemoglobin. Proc Natl Acad Sci U S A, 2000. 97(9): p. 4672-6.

41. Richardson, A.R., P.M. Dunman, and F.C. Fang, The nitrosative stress response of Staphylococcus aureus is required for resistance to innate immunity. Mol Microbiol, 2006. 61(4): p. 927-39.

42. Arai, H., et al., Transcriptional regulation of the flavohemoglobin gene for aerobic nitric oxide detoxification by the second nitric oxide-responsive regulator of Pseudomonas aeruginosa. J Bacteriol, 2005. 187(12): p. 3960-8.

43. de Jesus-Berrios, M., et al., Enzymes that counteract nitrosative stress promote fungal virulence. Curr Biol, 2003. 13(22): p. 1963-8.

44. Bang, I.S., et al., Maintenance of nitric oxide and redox homeostasis by the salmonella flavohemoglobin hmp. J Biol Chem, 2006. 281(38): p. 28039-47.

Page 184: thesis mtf 1 - Duke University

169

45. Jensen, D.E., G.K. Belka, and G.C. Du Bois, S-Nitrosoglutathione is a substrate for rat alcohol dehydrogenase class III isoenzyme. Biochem J, 1998. 331 ( Pt 2): p. 659-68.

46. Liu, L., et al., A metabolic enzyme for S-nitrosothiol conserved from bacteria to humans. Nature, 2001. 410(6827): p. 490-4.

47. Liu, L., et al., Essential roles of S-nitrosothiols in vascular homeostasis and endotoxic shock. Cell, 2004. 116(4): p. 617-28.

48. Liu, L., et al., Nitrosative stress: protection by glutathione-dependent formaldehyde dehydrogenase. Redox Rep, 2001. 6(4): p. 209-10.

49. Bateman, R.L., et al., Human carbonyl reductase 1 is an S-nitrosoglutathione reductase. J Biol Chem, 2008. 283(51): p. 35756-62.

50. Jaffrey, S.R., et al., Protein S-nitrosylation: a physiological signal for neuronal nitric oxide. Nat Cell Biol, 2001. 3(2): p. 193-7.

51. Chung, K.K., et al., S-nitrosylation of parkin regulates ubiquitination and compromises parkin's protective function. Science, 2004. 304(5675): p. 1328-31.

52. Uehara, T., et al., S-nitrosylated protein-disulphide isomerase links protein misfolding to neurodegeneration. Nature, 2006. 441(7092): p. 513-7.

53. Rizzo, M.A. and D.W. Piston, Regulation of beta cell glucokinase by S-nitrosylation and association with nitric oxide synthase. J Cell Biol, 2003. 161(2): p. 243-8.

54. Yasukawa, T., et al., S-nitrosylation-dependent inactivation of Akt/protein kinase B in insulin resistance. J Biol Chem, 2005. 280(9): p. 7511-8.

55. Wang, G., et al., Nitric oxide regulates endocytosis by S-nitrosylation of dynamin. Proc Natl Acad Sci U S A, 2006. 103(5): p. 1295-300.

Page 185: thesis mtf 1 - Duke University

170

56. Landino, L.M., et al., Ascorbic acid reduction of microtubule protein disulfides and its relevance to protein S-nitrosylation assays. Biochem Biophys Res Commun, 2006. 340(2): p. 347-52.

57. Dahm, C.C., K. Moore, and M.P. Murphy, Persistent S-nitrosation of complex I and other mitochondrial membrane proteins by S-nitrosothiols but not nitric oxide or peroxynitrite: implications for the interaction of nitric oxide with mitochondria. J Biol Chem, 2006. 281(15): p. 10056-65.

58. Kuncewicz, T., et al., Proteomic analysis of S-nitrosylated proteins in mesangial cells. Mol Cell Proteomics, 2003. 2(3): p. 156-63.

59. Huang, B. and C. Chen, An ascorbate-dependent artifact that interferes with the interpretation of the biotin switch assay. Free Radic Biol Med, 2006. 41(4): p. 562-7.

60. Njus, D. and P.M. Kelley, Vitamins C and E donate single hydrogen atoms in vivo. FEBS Lett, 1991. 284(2): p. 147-51.

61. May, J.M., J. Huang, and Z.C. Qu, Macrophage uptake and recycling of ascorbic acid: response to activation by lipopolysaccharide. Free Radic Biol Med, 2005. 39(11): p. 1449-59.

62. Winkler, B.S., S.M. Orselli, and T.S. Rex, The redox couple between glutathione and ascorbic acid: a chemical and physiological perspective. Free Radic Biol Med, 1994. 17(4): p. 333-49.

63. Bredt, D.S., C.D. Ferris, and S.H. Snyder, Nitric oxide synthase regulatory sites. Phosphorylation by cyclic AMP-dependent protein kinase, protein kinase C, and calcium/calmodulin protein kinase; identification of flavin and calmodulin binding sites. J Biol Chem, 1992. 267(16): p. 10976-81.

64. Foster, M.W. and J.S. Stamler, New insights into protein S-nitrosylation. Mitochondria as a model system. J Biol Chem, 2004. 279(24): p. 25891-7.

65. Becker, K. and R.H. Schirmer, 1,3-Bis(2-chloroethyl)-1-nitrosourea as thiol-carbamoylating agent in biological systems. Methods Enzymol, 1995. 251: p. 173-88.

Page 186: thesis mtf 1 - Duke University

171

66. Tonks, N.K., PTP1B: from the sidelines to the front lines! FEBS Lett, 2003. 546(1): p. 140-8.

67. Holmes, A.J. and D.L.H. Williams, Reaction of ascorbic acid with S-nitrosothiols: clear evidence for two distinct reaction pathways. J Chem Soc, Perkin Trans 2, 2000(8): p. 1639-1644.

68. Miles, A.M., et al., Photoinduced structural changes in the collagen/gelatin binding domain of fibronectin. Biochemistry, 1995. 34(21): p. 6941-6.

69. Permyakov, E.A., et al., Ultraviolet illumination-induced reduction of alpha-lactalbumin disulfide bridges. Proteins, 2003. 51(4): p. 498-503.

70. Fleming, J.E., et al., Interaction of ascorbic acid with disulfides. Zeitschrift fuer Naturforschung, C: Journal of Biosciences, 1983. 38C(9-10): p. 859-61.

71. D'Aquino, M., C. Dunster, and R.L. Willson, Vitamin A and glutathione-mediated free radical damage: competing reactions with polyunsaturated fatty acids and vitamin C. Biochem Biophys Res Commun, 1989. 161(3): p. 1199-203.

72. Chvanov, M., et al., Calcium-dependent release of NO from intracellular S-nitrosothiols. Embo J, 2006. 25(13): p. 3024-32.

73. Davisson, R.L., et al., Use-dependent loss of acetylcholine- and bradykinin-mediated vasodilation after nitric oxide synthase inhibition. Evidence for preformed stores of nitric oxide-containing factors in vascular endothelial cells. Hypertension, 1996. 28(3): p. 354-60.

74. Erwin, P.A., et al., Receptor-regulated dynamic S-nitrosylation of endothelial nitric-oxide synthase in vascular endothelial cells. J Biol Chem, 2005. 280(20): p. 19888-94.

75. Stamler, J.S., et al., Nitric oxide circulates in mammalian plasma primarily as an S-nitroso adduct of serum albumin. Proc Natl Acad Sci U S A, 1992. 89(16): p. 7674-7.

Page 187: thesis mtf 1 - Duke University

172

76. Beltran, B., et al., Oxidative stress and S-nitrosylation of proteins in cells. Br J Pharmacol, 2000. 129(5): p. 953-60.

77. Corrales, F.J., F. Ruiz, and J.M. Mato, In vivo regulation by glutathione of methionine adenosyltransferase S-nitrosylation in rat liver. J Hepatol, 1999. 31(5): p. 887-94.

78. Hao, G., L. Xie, and S.S. Gross, Argininosuccinate synthetase is reversibly inactivated by S-nitrosylation in vitro and in vivo. J Biol Chem, 2004. 279(35): p. 36192-200.

79. Barrett, D.M., et al., Inhibition of protein-tyrosine phosphatases by mild oxidative stresses is dependent on S-nitrosylation. J Biol Chem, 2005. 280(15): p. 14453-61.

80. Arrick, B.A. and C.F. Nathan, Glutathione metabolism as a determinant of therapeutic efficacy: a review. Cancer Res, 1984. 44(10): p. 4224-32.

81. Frischer, H. and T. Ahmad, Severe generalized glutathione reductase deficiency after antitumor chemotherapy with BCNU" [1,3-bis(chloroethyl)-1-nitrosourea]. J Lab Clin Med, 1977. 89(5): p. 1080-91.

82. Frischer, H., et al., Glutathione, cell proliferation, and 1,3-bis-(2-chloroethyl)-1-nitrosourea in K562 leukemia. J Clin Invest, 1993. 92(6): p. 2761-7.

83. Hara, M.R., et al., Neuroprotection by pharmacologic blockade of the GAPDH death cascade. Proc Natl Acad Sci U S A, 2006. 103(10): p. 3887-9.

84. Huang, Y., et al., S-nitrosylation of N-ethylmaleimide sensitive factor mediates surface expression of AMPA receptors. Neuron, 2005. 46(4): p. 533-40.

85. Kim, S.F., D.A. Huri, and S.H. Snyder, Inducible nitric oxide synthase binds, S-nitrosylates, and activates cyclooxygenase-2. Science, 2005. 310(5756): p. 1966-70.

86. Hou, Y., et al., Electrochemical studies of S-nitrosothiols. Bioorg Med Chem Lett, 1998. 8(21): p. 3065-70.

Page 188: thesis mtf 1 - Duke University

173

87. Williams, N.H. and J.K. Yandell, Outer-sphere electron-transfer reactions of ascorbate anions. Aust J Chem, 1982. 35(6): p. 1133-1144.

88. Koppenol, W.H., A thermodynamic appraisal of the radical sink hypothesis. Free Radic Biol Med, 1993. 14(1): p. 91-4.

89. Buettner, G.R., The pecking order of free radicals and antioxidants: lipid peroxidation, alpha-tocopherol, and ascorbate. Arch Biochem Biophys, 1993. 300(2): p. 535-43.

90. Walmsley, T.A., M.H. Abernethy, and H.P. Fitzgerald, Effect of daylight on the reaction of thiols with Ellman's reagent, 5,5'-dithiobis(2-nitrobenzoic acid). Clin Chem, 1987. 33(10): p. 1928-31.

91. Bartberger, M.D., et al., S-N dissociation energies of S-nitrosothiols: on the origins of nitrosothiol decomposition rates. J Am Chem Soc, 2001. 123(36): p. 8868-9.

92. Stamler, J.S. and E.J. Toone, The decomposition of thionitrites. Curr Opin Chem Biol, 2002. 6(6): p. 779-85.

93. Marzinzig, M., et al., Improved methods to measure end products of nitric oxide in biological fluids: nitrite, nitrate, and S-nitrosothiols. Nitric Oxide, 1997. 1(2): p. 177-89.

94. Moore, K.P. and A.R. Mani, Measurement of protein nitration and S-nitrosothiol formation in biology and medicine. Methods Enzymol, 2002. 359: p. 256-68.

95. Saville, B., A Scheme for the colorimetric determination of microgram amounts of thiols. Analyst, 1958. 83: p. 670-72.

96. Park, J.K. and P. Kostka, Fluorometric detection of biological S-nitrosothiols. Anal Biochem, 1997. 249(1): p. 61-6.

97. King, M., et al., Assessment of S-nitrosothiols on diaminofluorescein gels. Anal Biochem, 2005. 346(1): p. 69-76.

Page 189: thesis mtf 1 - Duke University

174

98. Kojima, H., et al., Detection and imaging of nitric oxide with novel fluorescent indicators: diaminofluoresceins. Anal Chem, 1998. 70(13): p. 2446-53.

99. Rodriguez, J., et al., Performance of diamino fluorophores for the localization of sources and targets of nitric oxide. Free Radic Biol Med, 2005. 38(3): p. 356-68.

100. Hausladen, A., et al., Assessment of nitric oxide signals by triiodide chemiluminescence. Proc Natl Acad Sci U S A, 2007. 104(7): p. 2157-62.

101. Doctor, A., et al., Hemoglobin conformation couples erythrocyte S-nitrosothiol content to O2 gradients. Proc Natl Acad Sci U S A, 2005. 102(16): p. 5709-14.

102. Palmer, L.A., et al., S-nitrosothiols signal hypoxia-mimetic vascular pathology. J Clin Invest, 2007. 117(9): p. 2592-601.

103. Palmer, L.A. and B. Gaston, S-nitrosothiol assays that avoid the use of iodine. Methods Enzymol, 2008. 440: p. 157-76.

104. Jia, L., et al., S-nitrosohaemoglobin: a dynamic activity of blood involved in vascular control. Nature, 1996. 380(6571): p. 221-6.

105. McMahon, T.J., et al., Nitric oxide in the human respiratory cycle. Nat Med, 2002. 8(7): p. 711-7.

106. Stamler, J.S., S-nitrosothiols in the blood: roles, amounts, and methods of analysis. Circ Res, 2004. 94(4): p. 414-7.

107. Mannick, J.B., et al., S-Nitrosylation of mitochondrial caspases. J Cell Biol, 2001. 154(6): p. 1111-6.

108. Haendeler, J., et al., Redox regulatory and anti-apoptotic functions of thioredoxin depend on S-nitrosylation at cysteine 69. Nat Cell Biol, 2002. 4(10): p. 743-9.

109. Park, H.S., et al., Nitric oxide negatively regulates c-Jun N-terminal kinase/stress-activated protein kinase by means of S-nitrosylation. Proc Natl Acad Sci U S A, 2000. 97(26): p. 14382-7.

Page 190: thesis mtf 1 - Duke University

175

110. Whalen, E.J., et al., Regulation of beta-adrenergic receptor signaling by S-nitrosylation of G-protein-coupled receptor kinase 2. Cell, 2007. 129(3): p. 511-22.

111. Sun, J., et al., Cysteine-3635 is responsible for skeletal muscle ryanodine receptor modulation by NO. Proc Natl Acad Sci U S A, 2001. 98(20): p. 11158-62.

112. Xu, L., et al., Activation of the cardiac calcium release channel (ryanodine receptor) by poly-S-nitrosylation. Science, 1998. 279(5348): p. 234-7.

113. Hausladen, A., et al., Nitrosative stress: activation of the transcription factor OxyR. Cell, 1996. 86(5): p. 719-29.

114. Leis, J.R., A. Rios, and L. Rodriguez-Sanchez, Reactivity of phenolic nucleophiles towards nitroso compounds. Part 2. Reaction with alkyl nitrite (O-nitroso compounds). J Chem Soc, Perkin Trans 2, 1998(12): p. 2729-33.

115. Munro, A.P. and D.L.H. Williams, Reactivity of nitrogen nucleophiles towards S-nitrosopenicillamine. J Chem Soc, Perkin Trans 2, 1999(10): p. 1989-93.

116. Forrester, M.T., M.W. Foster, and J.S. Stamler, Assessment and application of the biotin switch technique for examining protein S-nitrosylation under conditions of pharmacologically induced oxidative stress. J Biol Chem, 2007. 282(19): p. 13977-83.

117. Eu, J.P., et al., An apoptotic model for nitrosative stress. Biochemistry, 2000. 39(5): p. 1040-7.

118. Stuehr, D.J. and M.A. Marletta, Mammalian nitrate biosynthesis: mouse macrophages produce nitrite and nitrate in response to Escherichia coli lipopolysaccharide. Proc Natl Acad Sci U S A, 1985. 82(22): p. 7738-42.

119. Rodriguez, J., et al., Chemical nature of nitric oxide storage forms in rat vascular tissue. Proc Natl Acad Sci U S A, 2003. 100(1): p. 336-41.

Page 191: thesis mtf 1 - Duke University

176

120. Zhelyaskov, V.R., K.R. Gee, and D.W. Godwin, Control of NO concentration in solutions of nitrosothiol compounds by light. Photochem Photobiol, 1998. 67(3): p. 282-8.

121. Mayer, B., et al., Peroxynitrite-induced accumulation of cyclic GMP in endothelial cells and stimulation of purified soluble guanylyl cyclase. Dependence on glutathione and possible role of S-nitrosation. J Biol Chem, 1995. 270(29): p. 17355-60.

122. van der Vliet, A., et al., Formation of S-nitrosothiols via direct nucleophilic nitrosation of thiols by peroxynitrite with elimination of hydrogen peroxide. J Biol Chem, 1998. 273(46): p. 30255-62.

123. Schrammel, A., et al., S-nitrosation of glutathione by nitric oxide, peroxynitrite, and (*)NO/O(2)(*-). Free Radic Biol Med, 2003. 34(8): p. 1078-88.

124. Giustarini, D., et al., Is ascorbate able to reduce disulfide bridges? A cautionary note. Nitric Oxide, 2008. 19: p. (In Press).

125. Riddles, P.W., R.L. Blakeley, and B. Zerner, Reassessment of Ellman's reagent. Methods Enzymol, 1983. 91: p. 49-60.

126. May, J.M., et al., Ascorbate uptake and antioxidant function in peritoneal macrophages. Arch Biochem Biophys, 2005. 440(2): p. 165-72.

127. Winkler, B.S., Unequivocal evidence in support of the nonenzymatic redox coupling between glutathione/glutathione disulfide and ascorbic acid/dehydroascorbic acid. Biochim Biophys Acta, 1992. 1117(3): p. 287-90.

128. Nishikimi, M. and T. Ozawa, Stabilization of ascorbate solution by chelating agents that block redox cycling of metal ions. Biochem Int, 1987. 14(1): p. 111-7.

129. Oikawa, S. and S. Kawanishi, Distinct mechanisms of site-specific DNA damage induced by endogenous reductants in the presence of iron(III) and copper(II). Biochim Biophys Acta, 1998. 1399(1): p. 19-30.

Page 192: thesis mtf 1 - Duke University

177

130. Satoh, K., et al., Effect of metals and their antagonists on the radical intensity and cytotoxicity of ascorbates. Anticancer Res, 1997. 17(5A): p. 3355-60.

131. Saxena, P., et al., Transition metal-catalyzed oxidation of ascorbate in human cataract extracts: possible role of advanced glycation end products. Invest Ophthalmol Vis Sci, 2000. 41(6): p. 1473-81.

132. Wang, X., et al., Copper dependence of the biotin switch assay: modified assay for measuring cellular and blood nitrosated proteins. Free Radic Biol Med, 2008. 44(7): p. 1362-72.

133. Landgraf, W., et al., Oxidation of cysteines activates cGMP-dependent protein kinase. J Biol Chem, 1991. 266(25): p. 16305-11.

134. Schoneich, C., Selective Cu2+/ascorbate-dependent oxidation of alzheimer's disease beta-amyloid peptides. Ann N Y Acad Sci, 2004. 1012: p. 164-70.

135. Stohs, S.J. and D. Bagchi, Oxidative mechanisms in the toxicity of metal ions. Free Radic Biol Med, 1995. 18(2): p. 321-36.

136. Chevion, M., A site-specific mechanism for free radical induced biological damage: the essential role of redox-active transition metals. Free Radic Biol Med, 1988. 5(1): p. 27-37.

137. Gutierrez-Correa, J. and A.O. Stoppani, Inactivation of yeast glutathione reductase by Fenton systems: effect of metal chelators, catecholamines and thiol compounds. Free Radic Res, 1997. 27(6): p. 543-55.

138. Shapiro, M.P., B. Setlow, and P. Setlow, Killing of Bacillus subtilis spores by a modified Fenton reagent containing CuCl2 and ascorbic acid. Appl Environ Microbiol, 2004. 70(4): p. 2535-9.

139. Clarkson, T.W. and L. Magos, Studies on the binding of mercury in tissue homogenates. Biochem J, 1966. 99(1): p. 62-70.

140. Raftery, M.J., Enrichment by organomercurial agarose and identification of cys-containing peptides from yeast cell lysates. Anal Chem, 2008. 80(9): p. 3334-41.

Page 193: thesis mtf 1 - Duke University

178

141. Zalups, R.K., Molecular interactions with mercury in the kidney. Pharmacol Rev, 2000. 52(1): p. 113-43.

142. Zhang, Y., et al., Characterization and application of the biotin-switch assay for the identification of S-nitrosated proteins. Free Radic Biol Med, 2005. 38(7): p. 874-81.

143. Greco, T.M., et al., Identification of S-nitrosylation motifs by site-specific mapping of the S-nitrosocysteine proteome in human vascular smooth muscle cells. Proc Natl Acad Sci U S A, 2006. 103(19): p. 7420-5.

144. Han, P. and C. Chen, Detergent-free biotin switch combined with liquid chromatography/tandem mass spectrometry in the analysis of S-nitrosylated proteins. Rapid Commun Mass Spectrom, 2008. 22(8): p. 1137-45.

145. Hao, G., et al., SNOSID, a proteomic method for identification of cysteine S-nitrosylation sites in complex protein mixtures. Proc Natl Acad Sci U S A, 2006. 103(4): p. 1012-7.

146. Camerini, S., et al., A novel approach to identify proteins modified by nitric oxide: the HIS-TAG switch method. J Proteome Res, 2007. 6(8): p. 3224-31.

147. Kettenhofen, N.J., et al., In-gel detection of S-nitrosated proteins using fluorescence methods. Methods Enzymol, 2008. 441: p. 53-71.

148. Kettenhofen, N.J., et al., Proteomic methods for analysis of S-nitrosation. J Chromatogr B Analyt Technol Biomed Life Sci, 2007. 851(1-2): p. 152-9.

149. Han, P., et al., On-gel fluorescent visualization and the site identification of S-nitrosylated proteins. Anal Biochem, 2008. 377(2): p. 150-5.

150. Friedman, D.B. and K.S. Lilley, Optimizing the difference gel electrophoresis (DIGE) technology. Methods Mol Biol, 2008. 428: p. 93-124.

151. Wu, T.L., Two-dimensional difference gel electrophoresis. Methods Mol Biol, 2006. 328: p. 71-95.

Page 194: thesis mtf 1 - Duke University

179

152. Rhee, K.Y., et al., S-nitroso proteome of Mycobacterium tuberculosis: Enzymes of intermediary metabolism and antioxidant defense. Proc Natl Acad Sci U S A, 2005. 102(2): p. 467-72.

153. Benhar, M., et al., Regulated protein denitrosylation by cytosolic and mitochondrial thioredoxins. Science, 2008. 320(5879): p. 1050-4.

154. Roth, A.F., et al., Global analysis of protein palmitoylation in yeast. Cell, 2006. 125(5): p. 1003-13.

155. Huang, L., et al., Structure of an E6AP-UbcH7 complex: insights into ubiquitination by the E2-E3 enzyme cascade. Science, 1999. 286(5443): p. 1321-6.

156. Whitcomb, E.A., et al., Novel Control of S-Phase of the Cell Cycle by Ubiquitin Conjugating Enzyme H7. Mol Biol Cell, 2008.

157. Matthiesen, R. and K.E. Mutenda, Introduction to proteomics. Methods Mol Biol, 2007. 367: p. 1-35.

158. Ranish, J.A., M. Brand, and R. Aebersold, Using stable isotope tagging and mass spectrometry to characterize protein complexes and to detect changes in their composition. Methods Mol Biol, 2007. 359: p. 17-35.

159. Tenson, T. and A. Mankin, Antibiotics and the ribosome. Mol Microbiol, 2006. 59(6): p. 1664-77.

160. Cukras, A.R., et al., Ribosomal proteins S12 and S13 function as control elements for translocation of the mRNA:tRNA complex. Mol Cell, 2003. 12(2): p. 321-8.

161. Phillips, R.S. and P.D. Gollnick, Evidence that cysteine 298 is in the active site of tryptophan indole-lyase. J Biol Chem, 1989. 264(18): p. 10627-32.

162. Kruse, T., et al., Dysfunctional MreB inhibits chromosome segregation in Escherichia coli. EMBO J, 2003. 22(19): p. 5283-92.

Page 195: thesis mtf 1 - Duke University

180

163. Jones, L.J., R. Carballido-Lopez, and J. Errington, Control of cell shape in bacteria: helical, actin-like filaments in Bacillus subtilis. Cell, 2001. 104(6): p. 913-22.

164. van den Ent, F., L.A. Amos, and J. Lowe, Prokaryotic origin of the actin cytoskeleton. Nature, 2001. 413(6851): p. 39-44.

165. Gitai, Z., et al., MreB actin-mediated segregation of a specific region of a bacterial chromosome. Cell, 2005. 120(3): p. 329-41.

166. Robertson, G.T., et al., A Novel indole compound that inhibits Pseudomonas aeruginosa growth by targeting MreB is a substrate for MexAB-OprM. J Bacteriol, 2007. 189(19): p. 6870-81.

167. Hoffmann, J., et al., TNFalpha and oxLDL reduce protein S-nitrosylation in endothelial cells. J Biol Chem, 2001. 276(44): p. 41383-7.

168. Stennicke, H.R. and G.S. Salvesen, Caspases: preparation and characterization. Methods, 1999. 17(4): p. 313-9.

169. Cook, J.A., et al., Convenient colorimetric and fluorometric assays for S-nitrosothiols. Anal Biochem, 1996. 238(2): p. 150-8.

170. Nalvarte, I., et al., Overexpression of enzymatically active human cytosolic and mitochondrial thioredoxin reductase in HEK-293 cells. Effect on cell growth and differentiation. J Biol Chem, 2004. 279(52): p. 54510-7.

171. Jeong, W., et al., Roles of TRP14, a thioredoxin-related protein in tumor necrosis factor-alpha signaling pathways. J Biol Chem, 2004. 279(5): p. 3151-9.

172. Toyoshima, F., T. Moriguchi, and E. Nishida, Fas induces cytoplasmic apoptotic responses and activation of the MKK7-JNK/SAPK and MKK6-p38 pathways independent of CPP32-like proteases. J Cell Biol, 1997. 139(4): p. 1005-15.

173. Jeong, W., et al., Identification and characterization of TRP14, a thioredoxin-related protein of 14 kDa. New insights into the specificity of thioredoxin function. J Biol Chem, 2004. 279(5): p. 3142-50.

Page 196: thesis mtf 1 - Duke University

181

174. Arnelle, D.R. and J.S. Stamler, NO+, NO, and NO- donation by S-nitrosothiols: implications for regulation of physiological functions by S-nitrosylation and acceleration of disulfide formation. Arch Biochem Biophys, 1995. 318(2): p. 279-85.

175. Nikitovic, D. and A. Holmgren, S-nitrosoglutathione is cleaved by the thioredoxin system with liberation of glutathione and redox regulating nitric oxide. J Biol Chem, 1996. 271(32): p. 19180-5.

176. Ravi, K., et al., S-nitrosylation of endothelial nitric oxide synthase is associated with monomerization and decreased enzyme activity. Proc Natl Acad Sci U S A, 2004. 101(8): p. 2619-24.

177. Sliskovic, I., A. Raturi, and B. Mutus, Characterization of the S-denitrosation activity of protein disulfide isomerase. J Biol Chem, 2005. 280(10): p. 8733-41.

178. Stoyanovsky, D.A., et al., Thioredoxin and lipoic acid catalyze the denitrosation of low molecular weight and protein S-nitrosothiols. J Am Chem Soc, 2005. 127(45): p. 15815-23.

179. Sengupta, R., et al., Thioredoxin catalyzes the denitrosation of low-molecular mass and protein S-nitrosothiols. Biochemistry, 2007. 46(28): p. 8472-83.

180. Mitchell, D.A., et al., Thioredoxin is required for S-nitrosation of procaspase-3 and the inhibition of apoptosis in Jurkat cells. Proc Natl Acad Sci U S A, 2007. 104(28): p. 11609-14.

181. Arner, E.S., M. Bjornstedt, and A. Holmgren, 1-Chloro-2,4-dinitrobenzene is an irreversible inhibitor of human thioredoxin reductase. Loss of thioredoxin disulfide reductase activity is accompanied by a large increase in NADPH oxidase activity. J Biol Chem, 1995. 270(8): p. 3479-82.

182. Gromer, S., et al., Human placenta thioredoxin reductase. Isolation of the selenoenzyme, steady state kinetics, and inhibition by therapeutic gold compounds. J Biol Chem, 1998. 273(32): p. 20096-101.

183. Lillig, C.H. and A. Holmgren, Thioredoxin and related molecules-from biology to health and disease. Antioxid Redox Signal, 2007. 9(1): p. 25-47.

Page 197: thesis mtf 1 - Duke University

182

184. Verdoucq, L., et al., In vivo characterization of a thioredoxin h target protein defines a new peroxiredoxin family. J Biol Chem, 1999. 274(28): p. 19714-22.

185. Lee, S.R., et al., Molecular cloning and characterization of a mitochondrial selenocysteine-containing thioredoxin reductase from rat liver. J Biol Chem, 1999. 274(8): p. 4722-34.

186. Pia Rigobello, M., et al., Gold complexes inhibit mitochondrial thioredoxin reductase: consequences on mitochondrial functions. J Inorg Biochem, 2004. 98(10): p. 1634-41.

187. Nagata, S., Fas ligand-induced apoptosis. Annu Rev Genet, 1999. 33: p. 29-55.

188. Foster, M.W., T.J. McMahon, and J.S. Stamler, S-nitrosylation in health and disease. Trends Mol Med, 2003. 9(4): p. 160-8.

189. Kokkola, T., et al., S-nitrosothiols modulate G protein-coupled receptor signaling in a reversible and highly receptor-specific manner. BMC Cell Biol, 2005. 6(1): p. 21.

190. Nozik-Grayck, E., et al., S-nitrosoglutathione inhibits alpha1-adrenergic receptor-mediated vasoconstriction and ligand binding in pulmonary artery. Am J Physiol Lung Cell Mol Physiol, 2006. 290(1): p. L136-43.

191. Chanvorachote, P., et al., Nitric oxide negatively regulates Fas CD95-induced apoptosis through inhibition of ubiquitin-proteasome-mediated degradation of FLICE inhibitory protein. J Biol Chem, 2005. 280(51): p. 42044-50.

192. Kim, Y.M., R.V. Talanian, and T.R. Billiar, Nitric oxide inhibits apoptosis by preventing increases in caspase-3-like activity via two distinct mechanisms. J Biol Chem, 1997. 272(49): p. 31138-48.

193. Lipton, S.A., et al., A redox-based mechanism for the neuroprotective and neurodestructive effects of nitric oxide and related nitroso-compounds. Nature, 1993. 364(6438): p. 626-32.

Page 198: thesis mtf 1 - Duke University

183

194. Mustafa, A.K., et al., Nitric oxide S-nitrosylates serine racemase, mediating feedback inhibition of D-serine formation. Proc Natl Acad Sci U S A, 2007. 104(8): p. 2950-5.

195. Takahashi, H., et al., Hypoxia enhances S-nitrosylation-mediated NMDA receptor inhibition via a thiol oxygen sensor motif. Neuron, 2007. 53(1): p. 53-64.

196. Lowenstein, C.J., Nitric oxide regulation of protein trafficking in the cardiovascular system. Cardiovasc Res, 2007. 75(2): p. 240-6.

197. Matsushita, K., et al., Nitric oxide regulates exocytosis by S-nitrosylation of N-ethylmaleimide-sensitive factor. Cell, 2003. 115(2): p. 139-50.

198. Morrell, C.N., et al., Regulation of platelet granule exocytosis by S-nitrosylation. Proc Natl Acad Sci U S A, 2005. 102(10): p. 3782-7.

199. Tian, J., et al., S-nitrosylation/activation of COX-2 mediates NMDA neurotoxicity. Proc Natl Acad Sci U S A, 2008.

200. Xu, L., et al., Activation of cytosolic phospholipase A2alpha through nitric oxide-induced S-nitrosylation. Involvement of inducible nitric-oxide synthase and cyclooxygenase-2. J Biol Chem, 2008. 283(6): p. 3077-87.

201. Azad, N., et al., S-nitrosylation of Bcl-2 inhibits its ubiquitin-proteasomal degradation. A novel antiapoptotic mechanism that suppresses apoptosis. J Biol Chem, 2006. 281(45): p. 34124-34.

202. Chanvorachote, P., et al., Nitric oxide regulates cell sensitivity to cisplatin-induced apoptosis through S-nitrosylation and inhibition of Bcl-2 ubiquitination. Cancer Res, 2006. 66(12): p. 6353-60.

203. Lim, K.H., et al., Tumour maintenance is mediated by eNOS. Nature, 2008. 452(7187): p. 646-9.

204. Thomsen, L.L., et al., Selective inhibition of inducible nitric oxide synthase inhibits tumor growth in vivo: studies with 1400W, a novel inhibitor. Cancer Res, 1997. 57(15): p. 3300-4.

Page 199: thesis mtf 1 - Duke University

184

205. Fang, J., et al., S-nitrosylation of peroxiredoxin 2 promotes oxidative stress-induced neuronal cell death in Parkinson's disease. Proc Natl Acad Sci U S A, 2007. 104(47): p. 18742-7.

206. Gu, Z., et al., S-nitrosylation of matrix metalloproteinases: signaling pathway to neuronal cell death. Science, 2002. 297(5584): p. 1186-90.

207. Yao, D., et al., Nitrosative stress linked to sporadic Parkinson's disease: S-nitrosylation of parkin regulates its E3 ubiquitin ligase activity. Proc Natl Acad Sci U S A, 2004. 101(29): p. 10810-4.

208. Durham, W.J., et al., RyR1 S-nitrosylation underlies environmental heat stroke and sudden death in Y522S RyR1 knockin mice. Cell, 2008. 133(1): p. 53-65.

209. Iwakiri, Y., et al., Nitric oxide synthase generates nitric oxide locally to regulate compartmentalized protein S-nitrosylation and protein trafficking. Proc Natl Acad Sci U S A, 2006. 103(52): p. 19777-82.

210. Park, J.W., Reaction of S-nitrosoglutathione with sulfhydryl groups in protein. Biochem Biophys Res Commun, 1988. 152(2): p. 916-20.

211. Tsikas, D., et al., Investigations of S-transnitrosylation reactions between low- and high-molecular-weight S-nitroso compounds and their thiols by high-performance liquid chromatography and gas chromatography-mass spectrometry. Anal Biochem, 1999. 270(2): p. 231-41.

212. Ozawa, K., et al., S-nitrosylation of beta-arrestin regulates beta-adrenergic receptor trafficking. Mol Cell, 2008. 31(3): p. 395-405.

213. Holmgren, A., Thioredoxin and glutaredoxin systems. J Biol Chem, 1989. 264(24): p. 13963-6.

214. Stevanin, T.M., et al., Flavohemoglobin Hmp protects Salmonella enterica serovar typhimurium from nitric oxide-related killing by human macrophages. Infect Immun, 2002. 70(8): p. 4399-405.

Page 200: thesis mtf 1 - Duke University

185

215. Chen, K.S. and H.F. DeLuca, Isolation and characterization of a novel cDNA from HL-60 cells treated with 1,25-dihydroxyvitamin D-3. Biochim Biophys Acta, 1994. 1219(1): p. 26-32.

216. Nishiyama, A., et al., Identification of thioredoxin-binding protein-2/vitamin D(3) up-regulated protein 1 as a negative regulator of thioredoxin function and expression. J Biol Chem, 1999. 274(31): p. 21645-50.

217. Junn, E., et al., Vitamin D3 up-regulated protein 1 mediates oxidative stress via suppressing the thioredoxin function. J Immunol, 2000. 164(12): p. 6287-95.

218. Schulze, P.C., et al., Vitamin D3-upregulated protein-1 (VDUP-1) regulates redox-dependent vascular smooth muscle cell proliferation through interaction with thioredoxin. Circ Res, 2002. 91(8): p. 689-95.

219. Wang, Y., G.W. De Keulenaer, and R.T. Lee, Vitamin D(3)-up-regulated protein-1 is a stress-responsive gene that regulates cardiomyocyte viability through interaction with thioredoxin. J Biol Chem, 2002. 277(29): p. 26496-500.

220. Chen, J., et al., Thioredoxin-interacting protein deficiency induces Akt/Bcl-xL signaling and pancreatic beta-cell mass and protects against diabetes. FASEB J, 2008. 22(10): p. 3581-94.

221. Chutkow, W.A., et al., Thioredoxin-interacting protein (Txnip) is a critical regulator of hepatic glucose production. J Biol Chem, 2008. 283(4): p. 2397-406.

222. Parikh, H., et al., TXNIP regulates peripheral glucose metabolism in humans. PLoS Med, 2007. 4(5): p. e158.

223. Schulze, P.C., et al., Hyperglycemia promotes oxidative stress through inhibition of thioredoxin function by thioredoxin-interacting protein. J Biol Chem, 2004. 279(29): p. 30369-74.

224. Schulze, P.C., et al., Nitric oxide-dependent suppression of thioredoxin-interacting protein expression enhances thioredoxin activity. Arterioscler Thromb Vasc Biol, 2006. 26(12): p. 2666-72.

Page 201: thesis mtf 1 - Duke University

186

225. Kurtz, S., et al., The SecA2 secretion factor of Mycobacterium tuberculosis promotes growth in macrophages and inhibits the host immune response. Infect Immun, 2006. 74(12): p. 6855-64.

226. Hwang, C.Y., et al., Thioredoxin modulates activator protein 1 (AP-1) activity and p27Kip1 degradation through direct interaction with Jab1. Oncogene, 2004. 23(55): p. 8868-75.

227. Lyons, C.R., G.J. Orloff, and J.M. Cunningham, Molecular cloning and functional expression of an inducible nitric oxide synthase from a murine macrophage cell line. J Biol Chem, 1992. 267(9): p. 6370-4.

228. Xie, Q.W., R. Whisnant, and C. Nathan, Promoter of the mouse gene encoding calcium-independent nitric oxide synthase confers inducibility by interferon gamma and bacterial lipopolysaccharide. J Exp Med, 1993. 177(6): p. 1779-84.

229. Schmidt, H.H., et al., Regulation and subcellular location of nitrogen oxide synthases in RAW264.7 macrophages. Mol Pharmacol, 1992. 41(4): p. 615-24.

230. Diet, A., et al., Regulation of peroxiredoxins by nitric oxide in immunostimulated macrophages. J Biol Chem, 2007. 282(50): p. 36199-205.

231. Chen, P.F., et al., Mutation of Glu-361 in human endothelial nitric-oxide synthase selectively abolishes L-arginine binding without perturbing the behavior of heme and other redox centers. J Biol Chem, 1997. 272(10): p. 6114-8.

232. Intaglietta, M., P.C. Johnson, and R.M. Winslow, Microvascular and tissue oxygen distribution. Cardiovasc Res, 1996. 32(4): p. 632-43.

233. Tsai, A.G., et al., Microvascular and tissue oxygen gradients in the rat mesentery. Proc Natl Acad Sci U S A, 1998. 95(12): p. 6590-5.

234. Kaelin, W.G., Jr. and P.J. Ratcliffe, Oxygen sensing by metazoans: the central role of the HIF hydroxylase pathway. Mol Cell, 2008. 30(4): p. 393-402.

Page 202: thesis mtf 1 - Duke University

187

235. Darley-Usmar, V.M., et al., The simultaneous generation of superoxide and nitric oxide can initiate lipid peroxidation in human low density lipoprotein. Free Radic Res Commun, 1992. 17(1): p. 9-20.

236. Espey, M.G., et al., Focusing of nitric oxide mediated nitrosation and oxidative nitrosylation as a consequence of reaction with superoxide. Proc Natl Acad Sci U S A, 2002. 99(17): p. 11127-32.

237. Wink, D.A., et al., Reaction kinetics for nitrosation of cysteine and glutathione in aerobic nitric oxide solutions at neutral pH. Insights into the fate and physiological effects of intermediates generated in the NO/O2 reaction. Chem Res Toxicol, 1994. 7(4): p. 519-25.

238. Boon, E.M. and M.A. Marletta, Ligand discrimination in soluble guanylate cyclase and the H-NOX family of heme sensor proteins. Curr Opin Chem Biol, 2005. 9(5): p. 441-6.

239. Friebe, A. and D. Koesling, Regulation of nitric oxide-sensitive guanylyl cyclase. Circ Res, 2003. 93(2): p. 96-105.

240. Garthwaite, J., et al., Potent and selective inhibition of nitric oxide-sensitive guanylyl cyclase by 1H-[1,2,4]oxadiazolo[4,3-a]quinoxalin-1-one. Mol Pharmacol, 1995. 48(2): p. 184-8.

241. Francis, S.H., et al., Relaxation of vascular and tracheal smooth muscle by cyclic nucleotide analogs that preferentially activate purified cGMP-dependent protein kinase. Mol Pharmacol, 1988. 34(4): p. 506-17.

242. Ludwig, D.L., et al., Cloning, genetic characterization, and chromosomal mapping of the mouse VDUP1 gene. Gene, 2001. 269(1-2): p. 103-12.

243. Patwari, P., et al., The interaction of thioredoxin with Txnip. Evidence for formation of a mixed disulfide by disulfide exchange. J Biol Chem, 2006. 281(31): p. 21884-91.

244. Hashemy, S.I. and A. Holmgren, Regulation of the catalytic activity and structure of human thioredoxin 1 via oxidation and S-nitrosylation of cysteine residues. J Biol Chem, 2008. 283(32): p. 21890-8.

Page 203: thesis mtf 1 - Duke University

188

245. Mitchell, D.A. and M.A. Marletta, Thioredoxin catalyzes the S-nitrosation of the caspase-3 active site cysteine. Nat Chem Biol, 2005. 1(3): p. 154-8.

246. Sawa, A., et al., Glyceraldehyde-3-phosphate dehydrogenase: nuclear translocation participates in neuronal and nonneuronal cell death. Proc Natl Acad Sci U S A, 1997. 94(21): p. 11669-74.

247. Dooley, C.T., et al., Imaging dynamic redox changes in mammalian cells with green fluorescent protein indicators. J Biol Chem, 2004. 279(21): p. 22284-93.

248. Ostergaard, H., C. Tachibana, and J.R. Winther, Monitoring disulfide bond formation in the eukaryotic cytosol. J Cell Biol, 2004. 166(3): p. 337-45.

249. Sen, N., et al., Nitric oxide-induced nuclear GAPDH activates p300/CBP and mediates apoptosis. Nat Cell Biol, 2008. 10(7): p. 866-73.

Page 204: thesis mtf 1 - Duke University

189

Biography

Michael Tcheupdjian Forrester

Born April 19, 1980 in Chicago, IL

Education

2002-2004, 2009-2010; M.D., School of Medicine, Duke University, Durham, NC

2004-2009; Ph.D., Department of Biochemistry, Duke University, Durham, NC

1998-2002; B.A., Chemistry, Northwestern University, Evanston, IL

1994-1998; William Fremd High School, Palatine, IL

Manuscripts and Publications

[1] Forrester, M.T.; Foster, M.W.; Benhar, M.; Stamler, J.S. Detection of protein S-nitrosylation with the biotin-switch technique. Free Radic Biol Med 46: 119-126; 2009. [2] Benhar, M.; Forrester, M.T.; Hess, D.T.; Stamler, J.S. Regulated protein denitrosylation by cytosolic and mitochondrial thioredoxins. Science 320:1050-1054; 2008. [3] Forrester, M.T.; Foster, M.W.; Stamler, J.S. Assessment and application of the biotin switch technique for examining protein S-nitrosylation under conditions of pharmacologically induced oxidative stress. J Biol Chem 282:13977-13983; 2007. [4] Benhar, M.; Forrester, M.T.; Stamler, J.S. Nitrosative stress in the ER: a new role for S-nitrosylation in neurodegenerative diseases. ACS Chem Biol 1:355-358; 2006. [5] Forrester, M.T.; Stamler, J.S. A classification scheme for redox-based modifications of proteins. Am J Respir Cell Mol Biol 36:135-137; 2007. [6] Gomez-Vidal, J.A.; Forrester, M.T.; Silverman, R.B. Mild and selective sodium azide mediated cleavage of p-nitrobenzoic esters. Org Lett 3:2477-2479; 2001.